Sie sind auf Seite 1von 7

Colloids and Surfaces A: Physicochem. Eng.

Aspects 279 (2006) 121127

Generation of CdSe and CdTe nanoparticles by laser ablation in liquids


Albert A. Ruth , John A. Young
Physics Department, National University of Ireland, University College Cork, Cork, Ireland Received 7 July 2005; received in revised form 13 December 2005; accepted 26 December 2005 Available online 30 January 2006

Abstract Approximately spherical nanoparticles of the IIVI semiconductor materials CdSe and CdTe have been produced successfully by laser ablation of the bulk material in several liquids. The non-stabilized suspensions of particles are characterized by absorption spectroscopy and transmission electron microscopy (TEM). The procedure is not strongly size-selective, radii of 7 3 nm were found for CdSe and CdTe by transmission electron microscopy. Acetonitrile stabilizes the particles for several days up to weeks. Prolonged irradiation leads effectively to a reduction in particles size, in which particle agglomeration may play an important role. Ablation in degassed liquids does not have a signicant effect on the absorption of the suspended particles. Photoluminescence of laser-ablated CdSe and CdTe particles was not observed. 2006 Elsevier B.V. All rights reserved.
Keywords: Semiconductor; Laser processing; Electronic materials; Nanomaterials; Optical materials and properties

1. Introduction The remarkable photophysical and electrical properties of semiconductor nanoparticles [1,2] have triggered an enormous interest in new synthetic procedures for their production. The vast majority of established high-yield methods to synthesize nearly monodisperse IIVI semiconductor quantum dot distributions is based on chemical approaches [310]. The most widely used synthetic routes to synthesize CdSe and CdTe nanostructures use aqueous solutions [11,12] in which various functional groups (e.g. polar and unpolar thiols, amines and others) are employed to stabilize the particles. Common non-aqueous synthesis of crystalline IIVI semiconductors involve the decomposition of organometallic precursors in (mixtures of) the high boiling solvent trioctylphosphine (TOP) and its oxide (TOPO). Some of the wet chemical routes are photo-assisted [13]. Also an electropulse method has been reported for the synthesis of CdTe nanoparticles [14]. Purely physical production methods which allow the synthesis of uncapped particles are limited to mechanical ball milling processes [1517] and laser ablation in liquids. Using laser ablation in liquids to generate noble metal nanoparticles in suspension is known since 1993 [1820] and has since been also applied to several non-metallic systems [21,22].

Also semiconductors were recently produced by means of laser ablation at the bulk material-liquid interface. ZnSe [23], CdS and As2 S3 [2326], as well as GaAs [27] were the subjects of these studies. For CdTe the production of thin lms by laser deposition has recently been reported [28]. Here, we report the synthesis of CdSe and CdTe nanoparticles by laser ablation of the bulk material in different solvents and their characterization by absorption spectroscopy and transmission electron microscopy (TEM). The properties of the resulting particle, which are not stabilized by binding groups other than the solvent, will be discussed. 2. Experiment 2.1. Laser ablation The straightforward experimental setup for the ablation experiment is schematically shown in Fig. 1. The laser light used for the ablation was the second harmonic wave (532 nm) of a high frequency Nd:YAG laser (Laser Design, Germany). The pulses were focussed onto a piece of bulk semiconductor, which was held inside a quartz cell with 1 cm pathlength containing 3.5 ml of a solvent (see Fig. 1). The position of the sample cell with respect to the incident laser beam could be accurately adjusted in three directions by means of micrometer screws. Alignment of the laser on the target was achieved by wearing appropriate safety glasses which completely attenuate

Corresponding author. Fax: +353 21 4276949. E-mail address: a.ruth@ucc.ie (A.A. Ruth).

0927-7757/$ see front matter 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.colsurfa.2005.12.049

122

A.A. Ruth, J.A. Young / Colloids and Surfaces A: Physicochem. Eng. Aspects 279 (2006) 121127

before the measurement. Degassing of acetonitrile was done by several freeze-pump-thaw cycles. Bulk CdSe (purchased from Alfa Aesar) was of hexagonal structure and consisted of 46.8% Cd and 53.0% Se. The CdTe (Alfa Aesar) was of cubic structure and consisted of 46.9% Cd and 52.9% Te (percentages according to the manufacturers datasheet). Small pieces of the semiconductor beads were broken off the bulk and exposed to the laser pulses.
Fig. 1. Laser ablation setup schematically.

2.4. Transmission electron microscope Samples of nanoparticles for the electron microscope were prepared by placing a drop of suspension of interest on a copper mesh coated with an amorphous carbon lm (Agar Scientic). The drop was dried with a infrared lamp (Philips, 250 W) until all the solvent had evaporated. This process was repeated three to four times. The TEM carbon grids were loaded into the sample holder of a Joel 2000 FX TEM microscope (BioSciences Institute, University College Cork). The images were obtained at an accelerating voltage of 200 kV. 3. Results and discussion 3.1. Quantum connement The absorption spectra of laser-ablated CdSe and CdTe in suspension are shown in Figs. 2 and 3 (solid lines). The specic ablation conditions are outlined in the gure captions. Unlike the bulk material of CdSe and CdTe, which have steep absorption edges at 710 nm and 840 nm [29,30], respectively, the laser-ablated CdSe and CdTe nanoparticles show an onset of the absorption slightly above the band gap energy of the bulk at 700 nm for CdSe and 800 nm for CdTe. The absorption increases monotonically with decreasing wavelength towards

the intense (Rayleigh) scattered laser light and only transmit the red-shifted Raman scattered photons (Stokes). In that way, the laser beam can be viewed as it propoagates inside the liquid. With the laser operating at minimal power, which is insufcient to ablate the material, the Raman Stokes emission of the solvent was used to accurately adjust the position of the target in the focus of the lens. For all experiments a lens with a focal length of 5 cm was used. The laser pulses had a duration of 600 ns and pulse energies between 1 and 7 mJ. The repetition rate of the laser was generally 50 Hz and the total ablation time per measurement was typically several minutes. The estimated photon uences varied between 5 and 30 MW cm2 (see also [40]). Since nonlinear refraction and thermal heating can signicantly affect the focussing conditions, the photon uence at the surface of the semiconductor can vary from solvent to solvent and it depends on the particle yield. The uence also changes during the ablation process. At the start of an ablation experiment before nanoparticles are formed at a signicant number density, the focus is usually tighter, as particle shielding, and refractive and thermal effects are less pronounced. Generally laser ablations were executed in air-saturated solvents which were not stirred during the ablation. 2.2. Absorption spectra Absorption spectra were taken with a Lambda7 Perkin-Elmer double beam spectrometer. Before each ablation reference spectra of the pure solvents were measured in order to ensure that no traces of impurities remained in the ablation cell. The reference spectrum was also always subtracted from the absorption spectra of the colloidal suspensions. 2.3. Solvents and sample preparation The following solvents were used in the ablation experiments: (1) acetonitrile (SigmaAldrich, HPLC, 99.9%); (2) acetone (Lab-Scan Analytical Sciences, Capillary GC, 99.9%); (3) butan-2-ol (BDH Chemicals, 99.0%); (4) ethanol (Merck, Gradient Grade, 99.9%); (5) methanol (Lab-Scan Analytical Sciences, Capillary GC, 99.9%); (6) triethyleneglycol (Merck, >98%); (7) water (doubly distilled, Microanalysis Lab. UCC). All solvents were used without further purication. The optical cuvettes (Hellma Ltd.) were thoroughly cleaned and dried between experiments and ushed with the solvent of choice

Fig. 2. Solid line (ablation experiment): absorption spectrum of CdSe nanoparticles generated by laser ablation for 3 min in methanol (laser repetition rate 50 Hz, 3 mJ per pulse at 532 nm). Dashed and dotted lines (after ablation): absorption spectrum of the colloidal suspension (which was generated by laser ablation) after prolonged exposure to laser pulses for the times indicated in the graph (laser repetition rate 1500 Hz, 1.8 mJ per pulse at 532 nm). All spectra are taken at room temperature and are corrected for solvent absorption in the near UV and scattering. Insert: the same spectra scaled to the solid line at 300 nm.

A.A. Ruth, J.A. Young / Colloids and Surfaces A: Physicochem. Eng. Aspects 279 (2006) 121127

123

The small blue-shift of the absorption onset of the laser-ablated particles in comparison to the absorption edge of the bulk indicates the presence of quantum connement. The formation of approximately spherical CdSe and CdTe nanoparticles during laser ablation was conrmed through TEM, see Fig. 4. Stripes on many particles suggest a crystalline structure. For both materials the particles size distribution was found to be not very uniform with average radii of r = 7 3 nm as demonstrated in Fig. 4. Considering the observed size distribution it is a priori not obvious whether the majority of particles is in the weak (or strong) quantum connement limit, which is dened by the excitons Bohr radius, aB being much smaller (or larger) than the dimension of the spherical particle, respectively. Therefore aB was estimated for heavy-hole (hh) or light-hole (lh) excitons for CdSe and CdTe using
Fig. 3. Solid line (ablation experiment): absorption spectrum of CdTe nanoparticles generated by laser ablation for 3 min in methanol (laser repetition rate 50 Hz, 3 mJ per pulse at 532 nm). Dashed and dotted lines (after ablation): absorption spectrum of the colloidal suspension (which was generated by laser ablation) after prolonged exposure to laser pulses for the times indicated in the graph (laser repetition rate 1500 Hz, 1.8 mJ per pulse at 532 nm). All spectra are taken at room temperature and are corrected for solvent absorption in the near UV and scattering. Insert: the same spectra scaled to the solid line at 320 nm.

aB =

2 40 r h , e2

(1)

the UV. This indicates a rather wide distribution of particle sizes in comparison to chemically synthesized nanocrystals, which usually exhibit a well dened onset of the absorption spectrum well in the visible region and an absorption maximum according to the average (capped) particle size (see, e.g. Fig. 3 in [3]).

where e is the elementary charge; r , the static dielectric constant of the semiconductor (r = 10.2 [31] for CdSe and r = 10.6 for CdTe [32]) and = [me mh /(me + mh )] is the excitons reduced mass, which depends on the effective masses, me , of the electron and, mh , of the hole. The effective mass of the hole can be approximated by mlh for lh- and mhh for hh-transitions. Using me = 0.119 m0 , mhh = 0.820 m0 and mlh = 0.262 m0 [33] for CdSe, where m0 is the free electron mass, results in lh c(CdSe) 6.6 nm for light-hole, and ahh (CdSe) 5.2 nm aB B lh (CdTe) for heavy-hole excitons. For CdTe the Bohr radii aB

Fig. 4. TEM pictures of CdSe and CdTe nanoparticles generated in acetone and butan-2-ol, respectively (left) and the corresponding size distributions (right).

124

A.A. Ruth, J.A. Young / Colloids and Surfaces A: Physicochem. Eng. Aspects 279 (2006) 121127

Table 1 Comparison of the particle radii r determined from TEM measurements, with average particle radii rsc and rwc (strong and weak connement, Eqs. (2) and (4), respectively) for light- and heavy excitons. Material CdSea CdSeb CdTea CdTeb R exptl. (nm) 73 2 1c 73 rsc lh (nm) 9.7 2.3 7.0 2.0 rsc hh (nm) 8.1 2.0 5.8 1.7 rwc lh (nm) 11.9 7.4 10.9 7.9 rwc hh (nm) 12.4 9.7 11.5 9.6 aB lh (nm) 6.6 6.6 8.2 8.2 aB hh (nm) 5.2 5.2 6.0 6.0 E (eV) 0.02 0.70 0.08 1.20 Absorption onset (nm) 700 500 800 450

The corresponding Bohr radii aB (Eq. (1)) and differences of the band gap a Immediately after ablation. b After prolonged laser irradiation assuming scenario (B). c Aged sample, absorption not shown in Fig. 2.

E with respect to the bulk material are also given.

hh (CdTe) 6.0 nm were estimated using the vallimit is not reached according to the estimated Bohr radii in 8.2 nm and aB Table 1, the use of Eq. (4) is not recommended in the present ues me = 0.110 m0 , mhh = 0.600 m0 and mlh = 0.180 m0 [34].1 The case, since Eq. (4) yielded the best results for r/aB > 4 in Ref. Bohr radii are comparable with the average particle size found in [36]. It should be emphasised that all values stated in Table 1 the TEM studies, i.e. aB r. This indicates the regime between are only estimates. the weak-connement limit (r aB ) and the strong connement When after the initial laser ablation the particle suspensions limit (r aB ). In both limits the shift of the absorption edge of were additionally exposed to focussed laser pulses, the colour the nanoparticulate semiconductor in comparison to the bulk of the suspensions changed from light grey to almost transparcan be used to estimate the average particle radius. In the strong ent. The corresponding changes of the absorption spectra for connement (sc) case the shift in the band gap energy, E, is both CdSe and CdTe particles are also shown in Figs. 2 and 3 given in Refs. [35,36] as a function of r, which can hence be (dashed and dotted lines). The exposure times and laser concalculated as (in SI units): 2 2 2 2 2 1 1.786e h 1.786e rsc = ( E 0.248Eryd )1 (2) + 4( E 0.248Eryd ) 2 40 r 40 r 2

where Eryd e4 = 2(40 r )2h 2 (3)

is the effective Rydberg energy of the bulk exciton in SI units. In the weak connement (wc) case, where the exciton translational mass (me + mh ) is relevant (centre of mass localization), the empirical equation rwc = h + aB 2(mh + me )( E + Eryd ) (4)

can be applied (see Eq. (28) in [36]) in order to estimate the average particle radius. In Eq. (4), is a function of = mh /me which is approximately of the order of unity. is interpolated according to values given in Ref. [36]. It is not clear a priori whether Eqs. (2) or (4) is more appropriate in the present case. We used an absorption onset of 700 nm for CdSe (see Fig. 2) and 800 nm for CdTe (Fig. 3), as well as band-gaps of 1.751 and 1.475 eV for CdSe and CdTe [37], respectively, to calculate the radii rwc and rsc according to Eqs. (2) and (4). The results are listed in Table 1 for the same values of the effective masses as above. The values in Table 1 show that the calculated radii using Eq. (2) are in better agreement with the observed radii in the TEM study than the values estimated using Eq. (4). Even though the strong connement
1

ditions are stated in the gure captions. As opposed to the ablation experiment we chose a much higher laser repetition rate for the irradiation experiment since the rate of change of the sample absorption was small. The fact that the laser energy per pulse is smaller at the high repetition rate did not seem to affect the outcome of these experiments. In some samples agglomeration of particles in the form of tiny akes appearing in the suspension was observed upon prolonged exposure to laser pulses. At rst inspection, the change of the spectra could therefore be based on gravitational precipitation of larger particles, which would explain the absorbance decreasing over the entire wavelength range of the spectrum with increasing exposure time. However, scaling the spectra of CdSe and CdTe in Figs. 2 and 3 (see inserts) to the original spectra (solid line) at 300 and 320 nm, respectively,2 show that the irradiated suspensions exhibit a stronger absorption in the UV relative to the visible. This implies an increased quantum connement of the absorbing species, hence a reduction of the average size of the remaining nanoparticles in suspension. Two mechanisms concerning the laser-particle interaction are conceivable on the basis of these ndings: (A) laser-mergence of nanoparticle and (B) laser-disruption of larger nanostructures.

Effective masses were calculated in Refs. [33,34] for the [1 0 0] direction.

2 The scaling point was chosen such that all spectra agreed very well over the entire visible range of the spectrum (not fully shown)the fact that the scaling point is already in the near UV demonstrates that the relative increase in the deep UV is signicant.

A.A. Ruth, J.A. Young / Colloids and Surfaces A: Physicochem. Eng. Aspects 279 (2006) 121127

125

(A) The fact that agglomeration of particles was observed upon laser irradiation in some samples corroborates the assumption that particlelaser interaction promotes the formation of larger structures, which are subject to gravitational precipitation. The decrease in the absorption of the suspensions (Figs. 2 and 3) is simply due to a size dependent decrease in the number of nanoparticles. Assuming heavy (=large) particles, which are formed during laser irradiation, precipitate faster than light (=small) particles, the absorption is expected to remain stronger in the UV than in the visible with increasing exposure time (see insert Figs. 2 and 3). (B) In samples where no agglomeration was observed, larger nanoparticles could simply be broken up into smaller fractions upon interaction with intense laser pulses. In this case, the number density of smaller particles increases at the expense of that of larger particles, which is in agreement with the increase of the absorption in the UV relative to the visible (inserts in Figs. 2 and 3). Despite an increase of the overall number of nanoparticles in suspension the overall decrease of the absorption in Figs. 2 and 3 is still feasible, since the accurate dependence of the absorption cross-section of bare laser-ablated particles on size is not known. Provided the overall amount of material in suspension remains constant (which is the case for the outlined mechanism), the shift of the absorption onset of the semiconductor nanoparticles in suspension can be interpreted as being due to a larger band-gap of smaller nanoparticles being formed upon laser light exposure. For CdSe in acetonitrile the shift from originally 700 to 500 nm (solid and dotted lines in Fig. 2) corresponds to an increase of the band-gap of 0.7 eV. For CdTe in methanol the shift of the absorption onset from 800 to 450 nm corresponds to even 1.2 eV if the same scenario is valid. Using these shifts, average particle radii of the new suspension after laser irradiation were again estimated using Eqs. (2) and (4), the values are listed in Table 1. It is likely that both mechanisms (A) and (B) are active at the same time. The specic experimental conditions under which one of them dominates the other could not be established in this study. Simple sample ageing also led to a reduction in average particle size due to slow agglomeration and selective precipitation of larger particles. Suspensions after many weeks still exhibited a strongly UV biased absorption spectrum and TEM pictures revealed an average particle sizes of 2.2 0.8 nm, which compares well with values found for the strong connement limit in Table 1. 3.2. Dependence on the molecular environment The signicance of molecular oxygen for the ablation products was unclear at the outset of this study. All ablations were performed in air-saturated solvents. In order to acquire more information on whether (partially) oxidized particles are formed during laser ablation, experiments were performed using CdSe in degassed acetonitrile which was kept under vacuum. The expectation was to nd a difference in the absorption spectra of the

Fig. 5. Solid line (ablation experiment): absorption spectrum of CdSe nanoparticles generated by laser ablation for 15 min in acetonitrile (laser repetition rate 50 Hz, 3 mJ per pulse at 532 nm). Dashed and dotted lines (after ablation): absorption spectrum of the colloidal CdSe suspensions after prolonged exposure to laser pulses for the times indicated in the graph (laser repetition rate 1500 Hz, 1.8 mJ per pulse at 532 nm). All spectra are taken at room temperature and are corrected for solvent absorption in the near UV and scattering.

particle suspensions after laser ablation, if oxygen reacts with the suspended semiconductor particles during or after formation. The ablation efciency turned out to be much lower in the degassed solvent due to the formation of bubbles during the ablation, which strongly affects the focussing conditions. However, the absorption spectrum of the laser-ablated CdSe particles in the degassed suspension did not change dramatically in comparison to the air-saturated ones, as can be seen in Fig. 5 (ablation conditions in the caption). If oxygen happens to play a role during or after particle formation, it cannot be determined easily by UVvis absorption spectroscopy. Prolonged irradiation of the particles did not lead to a severe change, as it was observed in air-saturated samples, which may indicate that oxygen plays a role for the aking mechanisms discussed above. We failed to observe a photoluminescence from CdSe and CdTe nanoparticle suspensions upon excitation at several different wavelengths between 250 and 500 nm. According to Ref. [38], the luminescence properties of CdSe can be enhanced by means of surface passivation with oxygen. Provided reaction with oxygen increases the particles luminescence yield, the lack of a luminescence serves as an argument that laserablated nanoparticles are not or only weakly oxidized. Recently, it was reported that femtosecond laser pulse illumination in air of CdSe nanocrystalline lms on silicon enhanced the photoluminescence of such a system [39]. The prolonged irradiation of laser- ablated CdSe and CdTe particles in air-saturated suspensions did not change their photoluminescent properties measurably. Laser ablation of CdSe and CdTe was also performed in different solvents and under slightly different ablation conditions. The resulting absorption spectra of the CdSe and CdTe nanoparticles did not vary dramatically in different solvents as shown in Figs. 6 and 7. All CdSe and CdTe absorption spectra were normalized at 350 and 400 nm, respectively, for better comparison,

126

A.A. Ruth, J.A. Young / Colloids and Surfaces A: Physicochem. Eng. Aspects 279 (2006) 121127

Fig. 6. Comparison of the absorption spectra of suspended CdSe nanoparticles generated by ablation in different solvents. All spectra are normalized at 350 nm for better comparison. Laser repetition rate for all ablations: 50 Hz. Laser pulse energies (mJ) and ablation times (min): acetonitrile 3.5, 3; acetone 1.0, 30; butan-2-ol 7.0, 30; ethanol 1.5, 60; water 3.0, 60 (out-of-focus).

as ablation efciency and duration varied for the different solvents. At rst glance, the different refractive indeces of the solvents imply slightly different focussing conditions for each solvent. However, the change of the focussing conditions during the ablation due to heating of the solvent in the ablation plume environment is so large that the initial difference of the refractive indeces of the solvents is insignicant for the course of the ablation. At the start of the ablation (and in principle even throughout the duration of the ablation) the semiconductorsolvent system is not in thermal equilibrium. The refractive index and therefore the focussing conditions change according to the gradient (dn/dT) which depends on the specic heat and heat conductivity of the solvent and the target species. In Ref. [40], it was shown that the ablation of a silver target was largely indepen-

dent of the density, viscosity and the refractive index (at constant temperature) of a solvent. The laser uence did not appear to have a dramatic effect on the colloids too, which is demonstrated in the insert of Fig. 6. CdSe was ablated in acetonitrile (a) in the focus of the lens (high uence) and (b) 1 cm before the focal plane of the lens (focal length 5 cm). The two resulting spectra are virtually identical if normalized at 350 nm. For a lower photon uence, a longer ablation time is needed to obtain the same optical density of the resulting particle suspension. The semiconductor nanoparticles exhibited different stability in the different solvents. Suspensions of CdSe and CdTe in acetonitrile appeared to be most stable with precipitation times of many weeks. Acetone also stabilized both semiconductor nanoparticles rather well. In alcohols (methanol, butan-2-ol, ethanol, and also in triethyleneglycol) the particles fell out of suspension more rapidly, typically within days (in the case of CdSe in methanol even within a few hours). Even though CdSe also precipitated in ethanol, aged samples retained a strong UV absorption, and very small particles (r 14 nm) remained in suspension approximately 16 weeks after the initial ablation (see above). As evident in Figs. 6 and 7, water seems to have a signicantly different effect on the ablation than any other solvent. Water does not stabilize either semiconductor, CdSe and CdTe fell out of water suspensions within hours. This behaviour is quite different from laser-ablated noble metal nanoparticles, for which water is by far the best solvent for laser ablation at the metalliquid interface [40]. The reason for this may be that the focussing conditions in water change less during ablation than for the other solvents. The heat capacity of water is approximately twice as large as for the other solvents, its heat conductivity is approximately three times larger (causing efcient heat dissipation) and its refractive index temperature gradient (dn/dT) is approximately four to ve times smaller. The combination of a large heat capacity and conductivity slow the heating of water and the small change in the refractive index with increasing temperature indicate that the focussing conditions are less prone to change due to thermal effects during the ablation in water. This may explain the unusual results for semiconductorwater suspensions in Figs. 6 and 7. 4. Conclusions We demonstrated that approximately spherical CdSe and CdTe nanoparticles can be produced by laser ablation of the bulk material in solution. The particles are not highly monodispersed with radii of (7 3) nm as shown by TEM. The radii were estimated on basis of a pseudo-potential calculation in the weak and strong connement regime. The size distribution could be shifted by prolonged laser irradiation of suspensions after the initial ablation; two mechanisms for the size reduction were suggested. The absorption behaviour of nanoparticle suspension is not strongly dependent upon the presence of oxygen during and after ablation. The ablation is not strongly solvent dependent, with the exception of water. A reproducible photoluminescence from laser-ablated CdSe and CdTe particles in suspensions was not observed.

Fig. 7. Comparison of the absorption spectra of suspended CdTe nanoparticles generated by ablation in different solvents. All spectra are normalized at 400 nm for better comparison. Laser repetition rate for all ablations: 50 Hz. Laser pulse energies (mJ) and ablation times (min): Methanol 1.8, 15; acetone 7.0, 2; butan2-ol 7.0, 2; triethyleneglycol 7.0, 2; water 7.0, 2.

A.A. Ruth, J.A. Young / Colloids and Surfaces A: Physicochem. Eng. Aspects 279 (2006) 121127

127

Acknowledgements We are most grateful to Prof. N. Ernsting from the Department of Physical Chemistry, Humboldt University Berlin, who made the high-frequency Nd:YAG laser available to us. We markedly thank the Advanced Materials Group of the Chemistry Department, University College Cork (Dr. J. Holmes, Prof. M. Morris and Mr. J. Tobin), for providing access to their electron microscopy facility. Support by the Irish HEA PRTLI3 programme (Size, Shape & Spatial Control of Nanoparticles, Rods & Wires-Project 5) is gratefully acknowledged. References
[1] A.P. Alivisatos, Science 271 (1996) 933937. [2] A.P. Alivisatos, J. Phys. Chem. 100 (1996) 1322613239. [3] C.B. Murray, D.J. Norris, M.G. Bawendi, J. Am. Chem. Soc. 115 (1993) 87068715. [4] A.L. Rogach, L. Katsikas, A. Kornowski, D.S. Su, A. Eychm uller, H. Weller, Ber. Bunsen-Ges. Phys. Chem. 100 (1996) 17721778. [5] M.A. Hines, P. Guyot-Sionnest, J. Phys. Chem. 100 (1996) 468471. [6] B.O. Dabbousi, J. Rodriguez-Viejo, F.V. Mikulec, J.R. Heine, H. Mattoussi, R. Ober, K.F. Jensen, M.G. Bawendi, J. Phys. Chem. B 101 (1997) 94639475. [7] D.V. Talapin, A.L. Rogach, A. Kornowski, M. Haase, H. Weller, Nano Lett. 1 (2001) 207211. [8] N. Gaponik, D.V. Talapin, A.L. Rogach, K. Hoppe, E.V. Shevchenko, A. Kornowski, A. Eychm uller, H. Weller, J. Phys. Chem. B 106 (2002) 71777185. [9] M. Epifani, C. Giannini, L. Manna, Mater. Lett. 58 (2004) 24292432. [10] J.H. Yoon, W.S. Chae, S.J. Im, Y.R. Kim, Mater. Lett. 59 (2005) 14301433. [11] A. Henglein, Chem. Rev. 89 (1989) 18611873. [12] H. Weller, H. Schmidt, U. Koch, A. Fojtik, S. Baral, A. Henglein, W. Kunath, K. Weiss, Chem. Phys. Lett. 124 (1986) 557560. [13] Y.-W. Lin, M.-M. Hsieh, C.-P. Liu, H.-T. Chang, Langmuir 21 (2005) 728734. [14] J. Loicq, Y. Renotte, J.-L. Delplancke, Y. Lion, New J. Phys. 6 (2004) 32. [15] G.L. Tan, U. Hommerich, D. Temple, N.Q. Wu, J.G. Zheng, G. Loutts, Scripta Mater. 48 (2003) 14691474. [16] A. Urbieta, P. Fernandez, J. Piqueras, J. Appl. Phys. 96 (2004) 22102213.

[17] G.L. Tan, Q. Yang, U. Hommerich, J.T. Seo, D. Temple, Opt. Mater. 27 (2004) 579584. [18] A. Henglein, Israel J. Chem. 33 (1993) 7788. [19] A. Fojtik, A. Henglein, Ber. Bunsen-Ges. Phys. Chem. 97 (1993) 252254. [20] J. Neddersen, G. Chumanov, T.M. Cotton, Appl. Spectrosc. 47 (1993) 19591964. [21] Y.-P. Lee, Y.-H. Liu, C.-S. Yeh, Phys. Chem. Chem. Phys. 1 (1999) 46814686. [22] S.R.J. Pearce, S.J. Henley, F. Claeyssens, P.W. May, K.R. Hallam, J.A. Smith, K.N. Rosser, Diamond Rel. Mater. 13 (2004) 661 665. [23] K.V. Anikin, N.N. Melnik, A.V. Simakin, G.A. Shafeev, V.V. Voronov, A.G. Vitukhnovsky, Chem. Phys. Lett. 366 (2002) 357 360. [24] R.A. Ganeev, A.I. Ryasnyansky, T. Usmanov, Opt. Quant. Electron. 35 (2003) 211219. [25] R.A. Ganeev, A.I. Ryasnyansky, R.I. Tugushev, T. Usmanov, J. Opt. A, Pure Appl. Opt. 5 (2003) 409417. [26] R.A. Ganeev, A.I. Ryasnyansky, Opt. Commun. 246 (2005) 163171. [27] R.A. Ganeev, M. Baba, A.I. Ryasnyansky, M. Suzuki, H. Kuroda, Appl. Phys. B 80 (2005) 595601. [28] S.K. Pandey, U. Tiwari, R. Raman, C. Prakash, V. Krishna, V. Dutta, K. Zimik, Thin Solid Films 473 (2005) 5457. [29] C. Baban, G.I. Rusu, P. Prepelita, J. Optoelectron. Adv. Mater. 7 (2005) 817821, and references therein. [30] P. Bhattacharya, D.U. Bose, Semicond. Sci. Technol. 6 (1991) 384387. [31] B.G. Streetman, Solid State Electronic Devices, third ed., Prentice-Hall, New Jersey, 1990. [32] M.V.R. Krishna, R.A. Friesner, J. Chem. Phys. 95 (1991) 8309 8322. [33] H. Fu, L.-W. Wang, A. Zunger, Phys. Rev. B 57 (1998) 99719987. [34] F. Long, W.E. Hagston, P. Harrison, T. Stirner, J. Appl. Phys. 82 (1997) 34143421. [35] L.E. Brus, J. Chem. Phys. 80 (1984) 44034409. [36] Y. Kayanuma, Phys. Rev. B 38 (1988) 97979805. [37] J. Singh, Physics of Semiconductors and their Heterostructures, McGraw-Hill, New York, 1993. [38] N. Myung, Y. Bae, A.J. Bard, Nano Lett. 3 (2003) 747749. [39] M. Simurda, P. Nemec, F. Troj anek, P. Mal y, Thin Solid Films 453454 (2004) 300303. [40] J.A. Young, K.T. Lynch, A.J. Walsh, A.A. Ruth, Generation of noble metal nanoparticles by laser ablation in liquids: the role of the molecular environment, in: W.J. Blau, D. Kennedy, J. Colreavy (Eds.), Opto-Ireland 2005: Nanotechnology and Nanophotonics, SPIE-Proceedings, vol. 5824, 2005, pp. 138148.

Das könnte Ihnen auch gefallen