Você está na página 1de 5

NANO LETTERS

Synthesis of Highly Luminescent GaSe Nanoparticles


V. Chikan and D. F. Kelley*
Department of Chemistry, Kansas State UniVersity, Manhattan Kansas 66506-3701
Received October 4, 2001; Revised Manuscript Received November 5, 2001

2002 Vol. 2, No. 2 141-145

ABSTRACT
The synthesis, characterization, and purification of GaSe nanoparticles are described in this paper. TEM images show that this synthesis produces GaSe nanoparticles in the size range of 26 nm. These particles may be size segregated by column chromatography or size selective precipitation, and relatively monodisperse nanoparticles are obtained. Electron diffraction results indicate that these particles have a twodimensional single tetralayer type structure. The particles have absorption onsets in the 360 to 450 nm region, with the smallest particles absorbing furthest to the blue. The particles are emissive, with emission quantum yields of about 10%.

Downloaded by NAT LIB UKRAINE on August 6, 2009 Published on December 27, 2001 on http://pubs.acs.org | doi: 10.1021/nl015641m

Introduction. There has been a growing interest in the spectroscopic and photophysical properties of semiconductor nanoparticles, also referred to as nanocrystals or quantum dots. Despite this interest, relatively few types of semiconductor nanoparticles have photophysical properties that are well understood. The synthesis and characterization of these nanoparticles can be an involved and complex task. Furthermore, surface chemistry plays an important role in determining the size distribution and shape of the nanoparticles as well as their chemical and photophysical properties. Nanoparticles of layered semiconductors have twodimensional crystal and electronic structures, which can be very different than the three-dimensional structures of most semiconductor nanoparticles. We have recently studied nanoparticles of several types of layered semiconductors, specifically MoS2, WS2, MoSe2, and WSe2.1-3 We found that these nanoparticles consist of single trilayer (e.g., S-MoS) structures. This was expected, based on elementary chemical bonding considerations. There are strong covalent bonds within the S-Mo-S sheets, while only weak van der Waals forces hold adjacent sheets together. Detailed spectroscopic studies revealed that MoS2 nanoparticles have highly anisotropic optical and quantum confinement properties.1 We also found that, like the bulk materials,4 these nanoparticles are extremely photostable. GaSe has a hexagonal layered structure5 consisting of SeGa-Ga-Se sheets. Between the sheets there are only weak van der Waals interactions, which determine the physical appearance of the bulk solid. There are three different crystal structures of GaSe (, , and ), differing in how the SeGa-Ga-Se layers stack on each other. GaSe is an indirect band gap semiconductor6,7 having a 2.11 eV direct band gap. The difference between the direct and indirect band gap is
* To whom correspondence should be addressed: dfkelley@ksu.edu. 10.1021/nl015641m CCC: $22.00 Published on Web 12/27/2001 2002 American Chemical Society

quite small, about 25 meV. There are two different types of excitons in GaSe: two- and three-dimensional. The twodimensional exciton is associated with the indirect transition and is characterized by the effective mass tensor having a very large value along the (unique) z axis. The threedimensional exciton has an approximately isotropic effective mass, with a Bohr radius of 31 . The direct band gap transition is at and is characterized by having an electronic nodal plane between the layers of gallium atoms.8 Band gap excitation thus corresponds to the production of a node in the electronic wave function at the plane midway between the planes of gallium atoms. Thus, photoexcitation results in considerable Ga-Ga, * character, which does little to weaken the Ga-Se bonds. This is an important point because it suggests that this excited state may be far less reactive (and hence more photostable) than the band gap states of other types of semiconductor nanoparticles involving metalchalcogenide antibonding character. This situation is common in layered semiconductors and explains why WSe2-type materials are so stable as photoelectrodes.4,9 The possibility of producing highly luminescent, intrinsically photostable semiconductor nanoparticles provides much of the motivation for the work presented here. Two other groups have demonstrated the possibility of making GaSe nanoparticles. Allakhverdiev et al.10 made the first colloidal GaSe samples in 1997 using ultrasonic treatment of bulk GaSe in methanol. Their results show a wide distribution of GaSe particle sizes, based on the absorption spectrum. Stoll et al.11 have made GaSe nanoparticles using Ga4Se4R4 cubanes12,13 by MOCVD. The nanoparticles are about 88 nm in diameter, showing nanowire type structures in TEM images. In this paper, we report the results of a high temperature inorganic synthesis of surface capped GaSe nanoparticles. This method produces strongly quantum confined, highly luminescent GaSe nanoparticles.

Experimental Section. GaSe nanoparticles are synthesized using a method that is a modification of the well-known hightemperature synthesis of CdSe and InP nanoparticles.14-18 The GaSe synthesis is based on the reaction of an organometallic (GaMe3) with trioctyl phosphine selenium in a hightemperature solution of trioctyl phosphine (TOP) and trioctyl phosphine oxide (TOPO). The following synthetic procedure is used. A solution of 15 g of TOPO and 5 mL of TOP is heated to 150 C overnight in nitrogen atmosphere. Commercial TOPO is typically wet, and this heating removes any water by reaction with TOP to form TOPO. Prior to making this solution, the TOP (technical grade from Aldrich) is vacuum distilled at 0.75 Torr, taking the fraction from 204 C to 235 C. A TOPSe solution made from of 12.5 mL of TOP with 1.579 g Se (99.999%) is then added to the mixture. The above TOP/TOPO/TOPSe reaction mixture is heated to 278 C. This is followed by the injection of 0.8 mL of GaMe3 dissolved in 7.5 mL of distilled TOP. Upon injection, the temperature drops to 254 C and after 10 min stabilizes at 266-268 C. The presence of nanoparticles is indicated by the appearance of a 400-450 nm shoulder in the absorption spectrum. After this shoulder is formed, the reaction mixture is cooled to room temperature to prevent further reaction. The reaction takes about 2 h. After synthesis and cooling, the reaction vessel may be opened to the air and the nanoparticle solution may be extracted with methanol. The partitioning of the nanoparticles between the polar (methanol) and the nonpolar (TOP) phases depends on the experimental conditions. Under rigorously dry and anaerobic conditions, the nanoparticles extract into the methanol phase. Trace quantities of water and/or air apparently neutralize the charge on the particles, with the result being that the nanoparticles stay in the nonpolar phase. The results reported here are on nanoparticles that remain in the nonpolar phase upon methanol extraction. The particles are capped with TOPO at this point. We presume that the TOPO binds to the gallium atoms exposed at the particle edges. The TOPO may be displaced by the addition of about 1 m% 1-hexadecylamine to the nanoparticle solution. Following treatment with alkylamines the particles are somewhat less susceptible to aggregation, suggesting that amines bind more tightly than does TOPO. The nonpolar phase may be further purified by column chromatography. This may be done with nanoparticles that either have or have not been treated with an alkylamine. In the results reported here, the nanoparticles were treated with 1-hexadecylamine. Chromatographic separation of the nanoparticles was accomplished using a silica gel column with a mobile phase of 50 v% of CH2Cl2 and cyclohexane. Alternatively, particles may be size selectively precipitated by the addition of octane. The results of these purification procedures are discussed below. Absorption and luminescence spectra were acquired with an HP 8452 diode array absorption spectrometer and a SPEX FluoroMax-2 spectrophotometer, respectively. The nanoparticle emission quantum yields were determined following 400 nm excitation by comparison of the nanoparticle emission intensities with those from coumarin 314 dissolved
142

Downloaded by NAT LIB UKRAINE on August 6, 2009 Published on December 27, 2001 on http://pubs.acs.org | doi: 10.1021/nl015641m

Figure 1. Evolution of absorption spectrum of GaSe nanoparticles taken every half hour during synthesis (solid curves). Also shown is the spectrum following methanol extraction (dashed curve).

in ethanol (em ) 0.68). The optical density of the coumarin standard is adjusted to match the optical density of nanoparticle solution at 400 nm. The integrated areas of emission spectra are then directly comparable to determine the emission quantum yield. The emission spectra of coumarin and of the nanoparticles solution are reasonably overlapped, making this direct comparison valid. In all cases, the size distributions of the different nanocluster samples are determined with transmission electron microscopy (TEM). This was done using a Philips CM-200 TEM, operating at 100 kV. A small fraction of the nanocluster solution (either acetonitrile or octane phase) is diluted by a factor of 5, and a drop of the diluted solution is spread over a copper grid (300 mesh size) supporting a thin film of amorphous carbon. To reduce the damage from the electron beam, the sample is cooled to liquid nitrogen temperature during data collection. Electron diffraction results were obtained with the same instrument and on the same samples. Results and Discussion. The critical parameter of this synthesis, regarding the average particle diameter and size distribution, is the reaction temperature. The optimal synthesis temperature is found to be 268 C shortly after the injection of the GaMe3/TOP solution. Higher temperatures yield polydisperse nanoparticles, and at lower temperatures the reaction proceeds very slowly. The progress of the reaction and the nanoparticle size distribution is monitored by transferring a small aliquot of the reaction mixture to a cuvette and recording the absorption spectrum. The time evolution of the absorption spectrum is shown in Figure 1. The spectrum has an onset in the 400-450 nm region. Two control experiments establish that this absorption is due to the formation of GaSe nanoparticles. Either the selenium metal or the GaMe3 reactant has been left out from the reaction mixture, keeping all other parameters constant. In neither case is an absorption shoulder in this spectral region formed. The nanoparticle solution is initially purified by repeated washing (extraction) with methanol. The absorption spectrum
Nano Lett., Vol. 2, No. 2, 2002

Figure 2. (a, left panel) TEM image of GaSe nanoparticles following purification by methanol extraction. (b, right panel) TEM image of GaSe nanoparticles following chromatographic purification. This image corresponds to the first fraction off the column.

Downloaded by NAT LIB UKRAINE on August 6, 2009 Published on December 27, 2001 on http://pubs.acs.org | doi: 10.1021/nl015641m

Figure 4. Electron diffraction pattern of GaSe nanoparticles. Also shown are the assignments of several of the diffraction rings. The predicted locations of the (2,0,2) and (1,1,4) reflections are also shown as dashed and dotted rings, respectively.

Figure 3. TEM image of aggregated GaSe nanoparticles formed by size selective precipitation. The particles form an ordered hexagonal array.

after this purification (dashed line) is also shown in Figure 1. Methanol extraction results in an absorption onset that is shifted somewhat further to the blue and is more pronounced. This suggests that the largest particles produced by this synthesis are charged and preferentially go into the polar phase. Figure 2a shows a TEM image of GaSe nanoparticles following extraction of the nanoparticle solution with methanol several times. The nanoparticle diameters range from about 2 to 6 nm, with an average size of about 4 nm. Chromatographic purification results in very monodisperse samples, with different fractions corresponding to different sizes. We find that the smallest particles are eluted first. For example, chromatographic purification taking the first major fraction yields samples in which the average particle size has decreased to 2.5 nm. The size distribution of these nanoparticles is also narrowed to about (0.5 nm, see Figure 2b. The nanoparticle size distribution may also be narrowed by size selective precipitation. Addition of octane to the
Nano Lett., Vol. 2, No. 2, 2002

methanol-washed nanoparticle sample results in the precipitation and aggregation of the largest particles. Figure 3 shows a TEM image of the precipitated particles. These particles form aggregates that show highly ordered, hexagonally packed monodisperse arrays. In this case, the average (precipitated) particle size is about 6.1 nm, including the TOPO capping layer. Electron diffraction data have been obtained on methanolwashed nanoparticle samples and compared with diffraction patterns obtained on bulk GaSe and with crystallographic simulations. Several of the reflections observed for bulk GaSe are observed in the nanoparticle diffraction pattern, indicating the presence of crystalline GaSe (Figure 4). It is also important to note that several of the intense reflections seen for bulk GaSe are missing in the nanoparticle results. Specifically, the (2,0,2) and (1,1,4) reflections are observed and calculated to be intense in bulk GaSe and in large threedimensional (multilayer) GaSe particles. The reflections are calculated to get broader and less intense as the thickness of the nanoparticle is decreased, and to be absent in the case of a single tetralayer nanoparticle. Figure 4 shows that these reflections are absent. On the basis of this comparison, we conclude that GaSe nanoparticles have a single tetralayer structure. This is exactly what one would expect, based on simple bonding considerations. GaSe is a layered material
143

Figure 5. Emission spectra of GaSe nanoparticles before chromatographic purification excited at 320, 350, 400, and 450 nm.
Downloaded by NAT LIB UKRAINE on August 6, 2009 Published on December 27, 2001 on http://pubs.acs.org | doi: 10.1021/nl015641m

Figure 6. Absorption and emission spectra (excited at 350 nm) of GaSe nanoparticles before (solid line) and following (dashed line) chromatographic purification. The spectrum corresponds to the first fraction off the column.

consisting of covalently bound Ga-Se-Se-Ga tetralayers. The only forces holding the tetralayers to each other are relatively weak van der Waals forces. In the case of small (a few nanometers) particles, these forces are weak, and one would expect that solvent interactions would cause particles consisting of several tetralayers to exfoliate into single tetralayer particles. The GaSe nanoparticles described above are strongly emissive. The emission spectra following excitation at several wavelengths are shown in Figure 5. When the excitation wavelength is 350 nm or greater, the emission spectra show relatively high quantum yields, about 10%. This emission could be from any of several different electronic states. Bulk GaSe is an indirect band gap semiconductor having an energy difference between the lowest direct and indirect transitions of 25 meV. Either or both of these transitions may be involved in the observed emission. The observed emission may also involve trapped electrons and/or holes. Assignment of the transition(s) involved in this emission will require detailed spectroscopic studies and will be reported in a later paper. The emission spectra of GaSe nanoparticles obtained following excitation between 320 and 450 nm and purification by methanol extraction are shown in Figure 5. The excitation wavelength dependence is associated with the different sizes of GaSe nanoparticles. Excitation on the red edge of the absorption onset excites only the largest particles, while further blue excitation is not size selective. As a result, the emission following red-edge excitation is narrower and shifted further to the red. The emission is the most intense following excitation at 400 nm, which is the peak of the GaSe nanoparticle optical density. All the above results support the assignment that the GaSe nanoparticles are associated with a 400-450 nm absorption onset. This absorption onset is shifted 5000 to 8000 cm-1 to the blue of the (588 nm) bulk GaSe absorption onset, in accord with quantum confinement theories of semiconductor nanoparticles.
144

Figure 7. Emission spectra of 2.5 nm chromatographically purified GaSe nanoparticles excited at different wavelengths. The emission spectra correspond to excitation wavelengths from 330 to 410 nm in 10 nm increments. The emission maxima shift to the red with further red excitation.

Chromatographic purification significantly narrows the emission and absorption spectra as shown in Figure 6. In this case, the absorption and emission spectra of the smallest GaSe nanoparticles (2.5 nm) are shown and compared to the spectra obtained from the methanol-washed sample. The absorption spectrum of smaller, chromatographically purified nanoparticles is more distinct and shifted to the blue. The most intense emission also occurs at a bluer wavelength than before chromatographic purification (405 nm vs 435 nm). Figure 6 also shows that the emission spectrum of the chromatographically purified sample is narrower than in the unpurified case. This is because the chromatographically purified sample is more monodisperse. Figure 7 shows the emission spectra as a function of excitation wavelength of the chromatographically purified sample. It is of interest to compare these spectra to those of an unpurified sample, presented in Figure 5. The more polydisperse (unpurified)
Nano Lett., Vol. 2, No. 2, 2002

sample exhibits a strong dependence of the emission maximum with excitation wavelength. Because the chromatographically purified sample is more monodisperse, the emission spectra are less dependent on the excitation wavelength. Conclusions. In this paper, successful high temperature inorganic synthesis of GaSe nanoparticles has been demonstrated. The resulting particles can be further purified by chromatographic separation or size selective precipitation. In both cases, it is possible to narrow the nanoparticle size distribution. Smaller particles absorb and emit further to the blue than do larger particles. The resulting particles are highly luminescent, with emission quantum yields of about 10%. Further study is needed to understand the photophysics of these particles, and these results will be published in a later paper. Acknowledgment. The authors wish to thank Dr. Christer B. Aakeroy for the electron diffraction simulations. This work was supported by a grant from the U.S. Department of Energy (Grant # DE-FG03-00ER15037). References
(1) Chikan, V.; Kelley, D. F. J. Phys. Chem. B, submitted. (2) Huang, J. M.; Kelley, D. F. Chem. Mater. 2000, 12, 2825.

Downloaded by NAT LIB UKRAINE on August 6, 2009 Published on December 27, 2001 on http://pubs.acs.org | doi: 10.1021/nl015641m

(3) Huang, J. M.; Laitinen, R.; Kelley, D. F. Phys. ReV. B 2000, 62, 10995. (4) Solar Energy ConVersion; Seraphin, B. O., Ed.; Springer-Verlag: Berlin, 1979; Vol. 31. (5) Levy, F. Crystallography and Crystal Chemistry of Materials with Layered Structures; Reidel: Holland, 1976. (6) Lee, P. A. Physics and chemistry of materials with layered crystal structures; D. Reidel: Dordrecht, 1976; Vol. 4. (7) Grasso, V. Electronic structure and electronic transitions in layered materials; Reidel: Dordrecht, 1986. (8) Mooser, E.; Schluter, M. NuoVo Cimento 1973, 18B, 164. (9) Coehoorn, R.; Haas, C.; Dijkstra, J.; Flipse, C. J. F.; deGroot, R. A.; Wold, A. Phys. ReV. B 1987, 35, 6195. (10) Allakhverdiev, K.; Hagen, J.; Salaeva, Z. Phys. Status Solidi 1997, 163, 121-127. (11) Stoll, S. L.; Gillan, E. G.; Barron, A. Chem. Vap. Deposition 1996, 2, 182-184. (12) Gillan, E. G.; Barron, A. R. Chem. Mater. 1997, 9, 3037. (13) Gillan, E. G.; Bott, S. G.; Barron, A. R. Chem. Mater. 1997, 9, 796806. (14) Murray, C. B.; Kagan, C. R.; Bawendi, M. G. Science 1995, 270, 1335. (15) Peng, Z. A.; Peng, X. J. Am. Chem. Soc. 2001, 123, 183. (16) Talapin, D. V.; Rogach, A. L.; Kornowski, A.; Haase, M.; Weller, H. Nano Lett. 2001, 1, 207. (17) Guzelian, A. A.; Katari, J. E. B.; Kadavanich, V.; Banin, U.; Hamad, K.; Juban, E.; Alivisatos, A. P.; Wolters, R. H.; Arnold, C. C.; Heath, J. R. J. Phys. Chem. 1996, 100, 7272. (18) Micic, O. I.; Cheong, H. M.; Fu, H.; Zunger, A.; Sprague, J. R.; Mascarenhas, A.; Nozik, A. J. J. Phys. Chem. B 1997, 101, 4904.

NL015641M

Nano Lett., Vol. 2, No. 2, 2002

145

Você também pode gostar