Understanding the role of surfactants on the preparation of ZnS nanocrystals

Similar documents
Research Article. Optical and photoluminescence properties of HMTA capped transition metals (Cu, Co and Mn) doped ZnS nanoparticles

Effect of Reaction Tempreture on ME Capped ZnS Nanoparticles: Structural and Optical Characterization

CHAPTER 3. OPTICAL STUDIES ON SnS NANOPARTICLES

International Journal of Scientific & Engineering Research, Volume 5, Issue 3, March-2014 ISSN

SYNTHESIS OF CADMIUM SULFIDE NANOSTRUCTURES BY NOVEL PRECURSOR

Structural and Optical Properties of High-Purity Cubic Phase ZnS Nanoparticles Prepared by Thermal Decomposition Route for Optoelectronic Applications

CHAPTER III. ORGANIC LIGANDS PASSIVATED ZnSe NANOSTRUCTURES AND FUNCTIONAL PROPERTIES

The characterization of MnO nanostructures synthesized using the chemical bath deposition method

STUDIES ON ZnS - CuS NANOPARTICLE SYSTEM.

Supporting Information

Fluorescent silver nanoparticles via exploding wire technique

CHAPTER 3. SURFACE PASSIVATED CdS NANOPARTICLES AND THEIR PROPERTIES

CHAPTER 3. FABRICATION TECHNOLOGIES OF CdSe/ZnS / Au NANOPARTICLES AND NANODEVICES. 3.1 THE SYNTHESIS OF Citrate-Capped Au NANOPARTICLES

Synthesis and characterization of silica titania core shell particles

Influence of Nonionic Surfactant Concentration on Physical Characteristics of Resorcinol-Formaldehyde Carbon Cryogel Microspheres

Supporting Information

International Journal of Pure and Applied Sciences and Technology

Novel fluorescent matrix embedded carbon quantum dots enrouting stable gold and silver hydrosols

Metal nanoparticle-doped coloured films on glass and polycarbonate substrates

Enhanced photocurrent of ZnO nanorods array sensitized with graphene. quantum dots

Supporting Information

Conference Paper Synthesis and Efficient Phase Transfer of CdSe Nanoparticles for Hybrid Solar Cell Applications

Supplementary Information for Self-assembled, monodispersed, flowerlike γ-alooh

Detection of toxic mercury ions using a ratiometric CdSe/ZnS nanocrystal sensor

Electronic Supplementary Information for the Manuscript

Supporting Information. CdS/mesoporous ZnS core/shell particles for efficient and stable photocatalytic hydrogen evolution under visible light

Supporting Information. Synthesis and Upconversion Luminescence of BaY 2

Optical Properties and Characteristics of the CdSe Nanoparticles Synthesized at Room Temperature

PREPARATION OF LUMINESCENT SILICON NANOPARTICLES BY PHOTOTHERMAL AEROSOL SYNTHESIS FOLLOWED BY ACID ETCHING

SYNTHESIS OF ZnO:Mn NANOPARTICLES, NANOBELTS AND NANORODS

Applied Surface Science

Synthesis of ternary chalcogenide colloidal nanocrystals in aqueous medium

Supporting Information

The CdS and CdMnS nanocrystals have been characterized using UV-visible spectroscopy, TEM, FTIR, Particle Size Measurement and Photoluminiscence.

Supporting Information:

CH676 Physical Chemistry: Principles and Applications. CH676 Physical Chemistry: Principles and Applications

Hydrothermal synthesis and characterization of undoped and Eu doped ZnGa 2 O 4 nanoparticles

INVESTIGATIONS OF Mn, Fe, Ni AND Pb DOPED

A project report on SYNTHESIS AND CHARACTERISATION OF COPPER NANOPARTICLE-GRAPHENE COMPOSITE. Submitted by Arun Kumar Yelshetty Roll no 410 CY 5066

Synthesis of Zinc Sulphide Nanoparticles

SUPPORTING INFORMATION

Supporting Information. Solution-Based Growth of Monodisperse Cube-Like BaTiO 3 Colloidal Nanocrystals

CHAPTER-2 SYNTHESIS AND CHARACTERIZATION OF MODIFIED METAL OXIDES AND METAL SULPHIDES NANOPARTICLES

Synthesis with Different Se Concentrations and Optical Studies of CdSe Quantum Dots via Inverse Micelle Technique

PREPARATION AND CHARACTERIZATION OF CdO/PVP NANOPARTICLES BY PRECIPITATION METHOD

CHAPTER 4. ROLE OF CAPPING AGENTS IN PROMOTING THE MORPHOLOGICAL, STRUCTURAL AND OPTICAL PROPERTIES OF CdS NANOSTRUCTURES

Supporting Information. Dai-Wen Pang,

SYNTHESIS AND PROCESSING OF METALLIC NANOMATERIALS USING CO 2 EXPANDED LIQUIDS AS A GREEN SOLVENT MEDIUM

Supporting Information

Use of Ionic Liquids in the Synthesis of Nanocrystals and Nanorods of Semiconducting Metal Chalcogenides

Solution reduction synthesis of amine terminated carbon quantum dots

Phase Behaviour of Microemulsion Systems Containing Tween-80 and Brij-35 as Surfactant

Core-shell 2 mesoporous nanocarriers for metal-enhanced fluorescence

Supplementary Information for

A novel Ag 3 AsO 4 visible-light-responsive photocatalyst: facile synthesis and exceptional photocatalytic performance

Characterization of chemically synthesized CdS nanoparticles

Supporting Information

College of Mechanical Engineering, Yangzhou University, Yangzhou , China; 2

Synthesis and Characterization of Iron-Oxide (Hematite) Nanocrystals. Z.H. Lee

Supplementary Information

Zinc-Blende CdS Nanocubes with Coordinated Facets for Photocatalytic Water Splitting

Effect of solvent-induced structural modifications on optical properties of CdS nanoparticles

Microwave Synthesis of Monodisperse TiO 2 Quantum Dots and Enhanced Visible-Light Photocatalytic Properties

Fabrication and Characterization of Nanometer-sized AgCl/PMMA Hybrid Materials

Preparation of Silver Nanoparticles and Their Characterization

CHAPTER IV SYNTHESIS AND CHARACTERIZATION OF METAL IONS DOPED ZINC SELENIDE NANOPARTICLES

Markus Niederberger Max Planck Institute of Colloids and Interfaces, Potsdam, Germany.

Photoluminescence characteristics of Ti 2 S 3 nanoparticles embedded in sol gel derived silica xerogel

Effect of Solvents on the Morphological and Optical Properties

Facile bulk production of highly blue fluorescent

Study on the preparation and formation mechanism of barium sulphate nanoparticles modified by different organic acids

Photocatalytic degradation of dyes over graphene-gold nanocomposites under visible light irradiation

Electronic Supplementary Information. Phase transformation of mesoporous calcium carbonate by mechanical stirring

STRUCTURAL AND OPTICAL CHARACTERIZATION OF PVA CAPPED CADMIUM SULPHIDE NANOPARTICLES BY CHEMICAL COPRECIPITATION METHOD

Studying the Chemical, Optical and Catalytic Properties of Noble Metal (Pt, Pd, Ag, Au)/Cu 2 O Core-Shell Nanostructures Grown via General Approach

Multiply twinned Pt Pd nanoicosahedrons as highly active electrocatalyst for methanol oxidation

Microwave assisted synthesis of CdS nanoparticles for structural and optical Characterization

Module 4: "Surface Thermodynamics" Lecture 21: "" The Lecture Contains: Effect of surfactant on interfacial tension. Objectives_template

Supplementary Figure 1: (a) Upconversion emission spectra of the NaYF 4 4 core shell shell nanoparticles as a function of Tm

Classification of emulsifiers and stabilizers

FACILE SYNTHESIS OF CdS NANOCRYSTALS USING THIOGLYCOLIC ACID AS A SULFUR SOURCE AND STABILIZER IN AQUEOUS SOLUTION

Urchin-like Ni-P microstructures: A facile synthesis, properties. and application in the fast removal of heavy-metal ions

Growth of silver nanocrystals on graphene by simultaneous reduction of graphene oxide and silver ions with a rapid and efficient one-step approach

One-pot Solvent-free Synthesis of Sodium Benzoate from the Oxidation of Benzyl Alcohol over Novel Efficient AuAg/TiO 2 Catalysts

Supporting information

LUMINESCENCE SPECTRA OF QUANTUM-SIZED CdS AND PbI 2 PARTICLES IN STATIC ELECTRIC FIELD

Supplementary Information

*blood and bones contain colloids. *milk is a good example of a colloidal dispersion.

SUPPLEMENTARY INFORMATION

Preparation of Silver Iodide Nanoparticles Using a Spinning Disk Reactor in a Continuous Mode

Label-Free Fluorimetric Detection of Histone Using Quaternized Carbon Dot-DNA Nanobiohybrid. Electronic Supplementary Information (ESI)

Supporting Information

Visible-light Driven Plasmonic Photocatalyst Helical Chiral TiO 2 Nanofibers

Preparation of monodisperse silica particles with controllable size and shape

Supplementary Information

Electronic Supplementary Information (ESI)

Semiconductor quantum dots

Photocatalytic degradation of 4-nitrophenol in aqueous N, S-codoped TiO 2 suspensions

Supporting Information

In Situ synthesis of architecture for Strong Light-Matter Interactions

Transcription:

Journal of Colloid and Interface Science 297 (2006) 271 275 www.elsevier.com/locate/jcis Understanding the role of surfactants on the preparation of ZnS nanocrystals Milan Kanti Naskar, Amitava Patra, Minati Chatterjee Sol Gel Division, Central Glass & Ceramic Research Institute, Jadavpur, Kolkata 700 032, India Received 5 July 2005; accepted 20 October 2005 Available online 21 November 2005 Abstract We have synthesized surface modified ZnS nanoparticles of size 2 3 nm using non-ionic surfactant-stabilized reverse emulsions. The non-ionic surfactants in the Span series, i.e. sorbitan monolaurate (Span 20) and sorbitan monooleate (Span 80) of hydrophilic lipophilic balance (HLB) values of 8.6 and 4.3, respectively, have been used for the stabilization of emulsions. The role of these surfactants in controlling the size and properties of the ZnS nanoparticles has been discussed. The triethylamine (TEA) has been proved to be the effective surface modifying (capping) agent for the preparation of free-standing ZnS nanoparticles. The Span 20 with the higher HLB value of 8.6 has been found to be highly suitable in synthesizing TEA-capped ZnS nanoparticles of smaller size and higher photophysical characteristics compared to that of the Span 80 of lower HLB value of 4.3. A mechanism for the formation of TEA-capped ZnS nanoparticles from the surfactant-stabilized reverse emulsions has been proposed. 2005 Elsevier Inc. All rights reserved. Keywords: Nanoparticles; ZnS; Emulsion; Span-80; Photoluminescence 1. Introduction Semiconductor nanocrystals have attracted a lot of interest in recent years owing to their special electronic and optical properties due to quantum confinement effect [1 5]. Various wet chemical methods have been developed for the synthesis of these nanoparticles [6 10]. Synthesis of these nanoparticles with narrow size distribution is an essential requirement for practical purposes. Further, the optical properties are strongly dependent on the technique adopted for the chemical synthesis of these nanoparticles. Among other methods, reverse emulsion method is a well-accepted method for the synthesis of these nanoparticles due to several advantages. In our previous study, we have already seen the effect of non-ionic surfactants in tailoring the size of the alumina microspheres. In the present study, we will examine the role of these surfactants in synthesizing ZnS nanoparticles and their effects on photophysical properties which is not yet well studied. Keeping the above points in view, in the present investigation we report a reverse emulsion * Corresponding authors. Fax: +91 33 24730957. E-mail addresses: apatra@cgcri.res.in (A. Patra), minati@cgcri.res.in (M. Chatterjee). technique for the synthesis of triethylamine (TEA)-capped ZnS nanoparticles. In the present investigation, the two immiscible liquids, i.e. for the w/o in oil type emulsions (reverse emulsion), i.e. dispersion of the water phase (aqueous Zn 2+ in this study) in the oil phase (cyclohexane) are used. It is known that thermodynamically, the increase in surface area ( A) of the dispersed phase associated with a free energy ( G) can be represented as [12] G = γ A, where γ is the interfacial tension. The equation indicates that a low interfacial tension favours droplet disruption. In fact, the high interfacial tension between the dispersed water phase and continuous oil phase is reduced by the addition of a system compatible amphiphilic surface active agent or surfactant (emulsifier in the present case), i.e. the molecule with a polar head and a non-polar tail [12]. The surfactant molecules orient themselves according to the polarities of the involved chemical constituents. Thus, due to the high polarity of water, the polar heads of the surfactant molecules at the water oil interface are oriented towards the water droplets and get adsorbed on their surfaces and prevent the coalescence of the droplets by steric hindrance occurring between two droplets (steric stabilization). (1) 0021-9797/$ see front matter 2005 Elsevier Inc. All rights reserved. doi:10.1016/j.jcis.2005.10.057

272 M.K. Naskar et al. / Journal of Colloid and Interface Science 297 (2006) 271 275 (a) (b) Fig. 1. Molecular structures of (a) Span 80 and (b) Span 20. Therefore, the emulsifier added to the system: (i) lowers the interfacial tension between two immiscible liquids followed by droplet disruption and (ii) prevents re-coalescence of droplets by adsorbing on their surfaces, thus producing a stable emulsion. In the present investigation, Span 80 and Span 20 stabilized reverse emulsions comprising aqueous Zn 2+ solutions and cyclohexane as the water and oil phases, respectively, were used for the synthesis of ZnS nanoparticles. The surfactants in the Span series, i.e. Span 80, and Span 20 in the present case, are basically the fatty acid esters of anhydro sorbitols which are good oil soluble emulsifying agents [12]. A high (hydrophilic lyophobic balance) HLB value of the surfactant indicates strongly hydrophilic character while a low value is an indication of a strong hydrophobic nature. Considering these points, non-ionic surfactants, i.e. Span 80 and Span 20 with HLB values of 4.3 and 8.6, respectively, were selected for the present study. Figs. 1a and 1b depict the molecular structures of the amphiphilic Span 80 and Span 20, respectively, [11] where the hydrophilic sorbitan group acts as a polar head and the hydrophobic fatty acid group acts as the non-polar tail. The effect of the type of the surfactants on the photophysical properties of the nanocrystals is discussed. A mechanism for the separation free-standing ZnS nanocrystals from the stabilized emulsion is suggested. 2. Experimental 2.1. Synthesis of TEA-capped ZnS nanoparticles In the present investigation, water-in-oil (w/o) type emulsions were used for the synthesis of ZnS nanoparticles. For such synthesis, zinc acetate, dihydrate (A.R., Glaxo, India) and thioacetamide [CH 3 (CS)NH 2 )] (G.R., Loba Chemie, India) were used as the starting materials. An aqueous solution of zinc acetate with Zn 2+ concentration of 0.05 M and thioacetamide of concentration of 2.5 M was prepared by dissolving the respective salts in deionized water. The two solutions were mixed under stirring in the 1:1 volume ratio so as to obtain a mixed homogeneous solution; this mixed solution was used as the water phase (w) of w/o type emulsions. An organic solvent, i.e. cyclohexane (G.R., Merck, India) of dielectric constant 2.042 [8,11] at 25 C acted as the oil phase (o). A support solvent was prepared by mixing 5 vol% of the surfactant, sorbitan monooleate (Span 80) with HLB value of 4.3 (Fluka Chemie AG, Switzerland) in the cyclohexane. For the preparation of the w/o type emulsions, the water phase was dispersed as droplets in the support solvent under a constant mechanical agitation of 500 rpm and kept under such condition for 10 min for equilibration. The volume ratio of the water phase: support solvent was kept constant at 1:4 in all the experiments. The stabilized emulsion was then placed in an oil bath preset at 80 ± 1 C. The emulsion was aged at 80 ± 1 C for 10 min under stirring (500 rpm) and then removed from the bath. The entire emulsion was added to a known volume of acetone (G.R., Merck, India) (volume ratio of emulsion:acetone = 1:3) when the ph of the system was found to be 6. At this stage, sedimentation of the nanoparticles did not occur. To prepare freestanding capped-zns nanoparticles, the ph of the above system was then increased to 8 by the addition of triethylamine (TEA) (G.R., Merck, India). Immediate flocculation of particles occurred which were collected by centrifugation at 6000 rpm. To remove last traces of adhered impurities, the particles were washed thrice with acetone, each time collecting the particles centrifugally as described above, followed by the dispersion in the washing solvents. The washed particles were dried at 60 C. Following the same procedure another experiment was conducted using the surfactant Span 20. 2.2. Characterization of the sulphide nanoparticles The ZnS nanoparticles were characterized by XRD (Model: Philips, 1730, USA) with Ni-filtered Cu-K α radiation, FTIR study (Model: Nicolet 5PC, USA), and TEM (Model: JEOL JEJ-200X, Japan). Fluorescence spectra of ZnS nanoparticles dispersed in ethanol were recorded with a spectrofluorimeter (Model: LS 55, Perkin Elmer, USA) in the wavelength range of 200 650 nm. Optical absorbance of the ZnS particles were recorded with an UV vis spectrophotometer (Model: Cary 50, Varian) in the range 200 500 nm. 3. Results and discussion 3.1. Formation of ZnS nanoparticles The metal ions in the emulsified aqueous microdroplets were precipitated as sulphides by H 2 S, in situ generated at 80 ± 1 C via thermal decomposition of the thioacetamide [8]. The particle size in the microreactor was tailored by controlling the reaction temperature, ageing time and the concentration of Zn 2+ ions in the aqueous solutions. Based on the experimental results, the ageing time of 10 min at the reaction temperature of 80 ± 1 C and the concentration of Zn 2+ ions of 0.05 M were found to be the optimum. The particle size was further tailored by using surfactants of two different HLB values, keeping other

M.K. Naskar et al. / Journal of Colloid and Interface Science 297 (2006) 271 275 273 Fig. 2. Proposed mechanism for the separation of TEA-capped ZnS nanoparticles. parameters unaltered. Considering the HLB values and structures of the surfactants (Fig. 1) used in the present study, it is to be noted that the HLB value, i.e. the hydrophilicity, decreases with the increase in the hydrophobic chain length. Further, an inverse relationship between the average particle size and the HLB values of the surfactants was reported earlier during the synthesis of alumina microspheres by the sol emulsion gel method [11]. This was explained to be due to the fact that an increase in the HLB value caused a gradual decrease in the water oil interfacial tension, thus facilitating droplet disruption and finally giving rise to particles of progressive decreasing size. In our study, free-standing ZnS nanoparticles were obtained in both the Span 80- and Span 20-stabilized reverse emulsions, by increasing ph to 8 by the addition of TEA. The proposed mechanism for the separation of ZnS nanoparticles in presence of TEA is presented in Fig. 2. At ph 6, before the addition of TEA, ZnS nanoparticles formed complex with oxygen atoms of the polar heads of the surfactant molecules and caused an increase in apparent solubility in the acetone medium [13]. In presence of TEA, at ph 8, the above-mentioned ZnS complex became destabilized due to the basic hydrolysis of the surfactant molecules (the fatty acid esters) attached to the nanoparticles and formed new complex (capped) with the nitrogen atom of TEA molecules. The solubility of TEA-capped ZnS nanoparticles decreased in the acetone medium, leading to the flocculation as free-standing particles. The basic hydrolysis of the surfactant molecules is considered to be consisted of the following essential steps: (i) The TEA molecule in presence of H 2 O form triethylammonium hydroxide (C 2 H 5 ) 3 N + H 2 O (C 2 H 5 ) 3 N + HOH. (ii) The triethylammonium hydroxide increases the ph of the system by liberating hydroxide (OH ) ions. (iii) The (OH ) ion attacks the carbon (C) atom of the ester carbonyl (>C=O) group in the surfactant molecule (sor- Fig. 3. XRD patterns of TEA-capped ZnS nanoparticles using (a) Span 20 and (b) Span 80. bitan monooleate as a typical example), causing hydrolysis into the sorbitan and oleic acid parts (Fig. 2). (iv) Hydrolysis of the sorbitan monooleate destabilizes the emulsion system followed by the flocculation of the generated particles. Formation of TEA-capped ZnS nanoparticles is confirmed by XRD and FTIR studies described in the subsequent sections. The effect of the nature of the surfactants and their HLB values on the characteristics of the nanoparticles has been examined by the UV vis absorption and photoluminescence studies. 3.2. XRD and FTIR studies Fig. 3 a and b represent the powder XRD patterns of the TEA-capped ZnS nanoparticles prepared by our method with surfactants Span 80 and Span 20, respectively. It is to be noted that in both the figures the peaks observed in the XRD patterns match well with those of the β-zns (cubic) reported in the JCPDS Powder Diffraction File No. 5-0566. Intensities of the three most important peaks of ZnS, namely 111, 220 and 311 reflections corresponding to 28.5, 47.6 and 56.4 respectively, do not deviate from the Powder Diffraction File intensities [14]. Broadening of the XRD peaks shows the formation of nanocrystals of ZnS. XRD patterns also indicate that the peaks in Fig. 3b are broader than those in Fig. 3a. Crystallite size of ZnS nanoparticles were calculated following the Scherrer s equation [8,14] using 111 reflection of the XRD patterns and are estimated to be 2.4 and 1.8 nm for Fig. 3, a and b respectively. XRD results, therefore, indicate the formation of smaller size ZnS particles from the Span 20-stabilized reverse emulsion compared to that obtained from the Span 80. The presence of TEA in the synthesized ZnS nanoparticles was examined by recording their FTIR spectra in the range 4000 400 cm 1. Fig. 4 depicts the spectra of TEA-capped ZnS nanoparticles derived from Span 80-stabilized emulsion as a typical example. From Fig. 4, it is to be noted that the absorption peaks at 3431 and 1631 cm 1 in the figure are assigned

274 M.K. Naskar et al. / Journal of Colloid and Interface Science 297 (2006) 271 275 Fig. 4. FTIR spectrum of TEA-capped ZnS nanoparticles. Fig. 6. UV vis absorption spectra of ZnS nanoparticles using (a) Span 80 and (b) Span 20. 3.4. UV vis absorption and PL studies (a) (b) Fig. 5. TEM micrographs of ZnS nanoparticles using (a) Span 80 and (b) Span 20. to water molecules and hydrogen bonded hydroxyl groups respectively [14,15]. Other peaks observed mainly at 2923, 2846, 1554, 1445, 1377 and 1062 cm 1 are due to C H bonding [14,15]. This is due to the formation of co-ordinate bond between the nitrogen atom of the TEA and Zn 2+ ions [14,15]. Therefore, from the FTIR results in the present study, capping of ZnS nanoparticles with TEA molecules is assured in our samples. The effect of particle size on the UV vis absorption and PL studies has been discussed below. 3.3. TEM study The TEM micrographs of the Span 80- and Span 20-derived TEA-capped nanoparticles are depicted in Figs. 5a and 5b respectively. In both the micrographs, the particles are monodispersed in nature and the size was estimated to be in the range 2 3 nm which corroborated with the XRD results. UV vis absorption spectra of TEA-capped ZnS nanoparticles obtained from Span 80- and Span 20-stabilized reverse emulsions have been presented in Fig. 6 a and b, respectively. Fig. 6a exhibits the absorption edge at 303 nm with a shoulder peak at 285 nm, while in Fig. 6b the absorption edge at 290 nm with a shoulder peak at 272 nm is noticed. Absorption edges in Fig. 6, a and b exhibit blue shifts in comparison to bulk ZnS (345 nm). This shift of the absorption edges to shorter wavelengths is explained due to the quantum confinement of ZnS nanoparticles [6,14]. The absorption edges of the nanocrystallites are also sharp, indicating that the synthesized particles have relatively narrow size distributions. Here, we calculate the cluster size of ZnS using the following equation [16,17] E = π 2 h 2 /2µr 2 1.78e 2 /ɛr, where E is blue shift band gap and µ is the reduced mass of electron and hole. 0.2 is taken as the electron rest mass m 0, r is the cluster size and ɛ is the dielectric constant. The first term indicates the confinement effect and the second term is the Coulomb term. In a strong confinement, as in the present case, the second term is small and can be neglected. Based on Eq. (2), the cluster sizes of Span 80- and Span 20-derived ZnS have been found to be 2.1 and 2.6 nm respectively, which is fairly consistent with those obtained from the XRD study. Therefore, from the optical absorption study we infer that the surfactant Span 20 with the higher HLB value of 8.6 successfully produced smaller ZnS nanoparticles compared to that obtained from the Span 80 of lower HLB value. The PL spectra of the Span 80- and Span 20-derived TEAcapped ZnS nanoparticles are presented in Fig. 7 a and b, respectively. The band edge emissions peaking at 379 and 373 nm are observed for the samples corresponding to Fig. 7 a and b, respectively. In both the spectra, PL peaks are markedly blue shifted (about 65 nm) relative to that of bulk ZnS (440 nm) and thus exhibiting quantum size effects [14,17]. The interesting point is that under identical conditions of synthesis, the Span 20-derived nanoparticles exhibit highly intense (about eight times) PL emission than that of the Span 80-derived nanoparticles. Further, the narrow emission band width of the (2)

M.K. Naskar et al. / Journal of Colloid and Interface Science 297 (2006) 271 275 275 We have reported a soft chemical method for in situ synthesis of monodispersed, TEA-capped ZnS nanoparticles using two different surfactant-stabilized w/o type reverse emulsions. Two types of non-ionic surfactants in the Span series, i.e. Span 80 and Span 20 with the HLB values of 4.3 and 8.6, respectively, have been used for the stabilization of emulsions in the present study. A mechanism for the generation of TEAcapped ZnS nanoparticles from stabilized emulsions has been suggested. The role of surfactants on the characteristics of ZnS nanoparticles has been discussed. XRD reveals that the synthesized nanoparticles exhibit zinc blende (cubic) structure. Crystallite size measured from XRD peaks proves the formation of smaller (1.8 nm) ZnS nanoparticles from the Span 20-stabilized emulsions compared to that (2.4 nm) of the Span 80. Formation of TEA-capped nanoparticles is confirmed by the FTIR studies. TEM micrographs exhibit well-defined, monodispersed ZnS nanoparticles of size 2 3 nm which corroborates with the XRD results. UV vis absorption and PL studies further prove that the nanoparticles with comparatively smaller size and narrow size distribution obtained from Span 20-stabilized reverse emulsion exhibited strong quantum size effect and highly intense PL characteristics compared to that of the Span 80-derived one. Acknowledgments Fig. 7. Photoluminescence spectra of ZnS nanoparticles using (a) Span 80 and (b) Span 20. same sample indicates a relatively narrow size distribution of the particles. The red shifting of the PL peak to 379 nm for the Span 80-derived ZnS nanoparticles compared to that of at 373 nm of the Span 20-derived particles, explains the formation of smaller particles in the Span 20-derived particles. The PL emission in both the samples was assigned to the interstitial sulphur lattice defects [14,18]. Therefore, PL studies further prove the formation of smaller sized ZnS nanoparticles from the Span 20-stabilized reverse emulsions compared to that from the Span 80. This happens because Span 80 with low HLB value of 4.3, produced large size microdroplets and caused more ZnS particle formation via precipitation which subsequently formed larger particles by Ostwald ripening during ageing at the reaction temperature compared to that of Span 20 of higher HLB value of 8.6. 4. Summary Authors thank Dr. H.S. Maiti, Director of CGCRI for his constant encouragement and active co-operation to carry out the work. This work is partially financial supported by the Department of Science & Technology (DST), New Delhi (Project No. III.5(1)/2000-ET). References [1] R.N. Bhargava, D. Gallagher, X. Hong, A. Nurmikko, Phys. Rev. Lett. 72 (1994) 416. [2] T. Schmidt, G. Muller, L. Spanhel, K. Kerkel, A. Forchel, Chem. Mater. 10 (1998) 65. [3] S. Sapra, A. Prakash, A. Ghangrekar, N. Periasamy, D.D. Sarma, J. Phys. Chem. B 109 (2005) 1663. [4] W. Chen, J.O. Malm, V. Zwiller, R. Wallenberg, J.O. Bovin, J. Appl. Phys. 89 (2001) 2671. [5] M.A. Hines, P.G. Sionnest, J. Phys. Chem. 100 (1996) 468. [6] T. Trindade, P. O Brien, Chem. Mater. 9 (1997) 523. [7] M. Azad Malik, P. O Brien, N. Revaprasadu, J. Mater. Chem. 11 (2001) 2382. [8] M. Chatterjee, A. Patra, J. Am. Ceram. Soc. 84 (2001) 44. [9] N. Arul Dhas, A. Gedanken, Appl. Phys. Lett. 72 (1998) 2514. [10] J. Nanda, S. Sapra, D.D. Srama, N. Chandrasekharan, G. Hodes, Chem. Mater. 12 (2000) 1018. [11] M. Chatterjee, M.K. Naskar, B. Siladitya, D. Ganguli, J. Mater. Res. 15 (2000) 176. [12] E. Dickinson, in: D.J. Wedlock (Ed.), Controlled Particle, Droplet and Bubble Formation, Butterworth Heinemann, Oxford, 1994, p. 191. [13] M. Chatterjee, M.K. Naskar, A. Patra, Indian Patent Application No. 1295/ DEL/01, dated 28/12/01. [14] G. Ghosh, M.K. Naskar, A. Patra, M. Chatterjee, Opt. Mater. (2005), submitted for publication. [15] H. Tang, M. Yan, H. Zhang, M. Xia, D. Yang, Mater. Lett. 59 (2005) 1024. [16] A.B. Yoffe, Adv. Phys. 42 (1993) 173. [17] J.X. Yin, J. Yan, Z. Zhi-Lin, Xu, Shao-Hong, J. Cryst. Growth 191 (1998) 692. [18] D. Denzler, M. Olschewski, K. Sattler, J. Appl. Phys. 84 (1998) 2841.