Reversal of Hall–Petch Effect in Structural Stability of PbTe

Nov 1, 2011 - State University of New York at Binghamton, Binghamton, New York 13902 ... from chemistry, physics, and materials science to geophysics...
0 downloads 0 Views 3MB Size
LETTER pubs.acs.org/NanoLett

Reversal of Hall Petch Effect in Structural Stability of PbTe Nanocrystals and Associated Variation of Phase Transformation Zewei Quan,† Yuxuan Wang,‡ In-Tae Bae,§ Welley Siu Loc,† Chenyu Wang,† Zhongwu Wang,*,|| and Jiye Fang*,†,‡ Department of Chemistry, ‡Materials Science and Engineering Program, and §Small Scale Systems Integration and Packaging Center, State University of New York at Binghamton, Binghamton, New York 13902, United States Cornell High Energy Synchrotron Source, Wilson Laboratory, Cornell University, Ithaca, New York 14853, United States

)



bS Supporting Information ABSTRACT: Using an in situ synchrotron X-ray diffraction technique, a pressureinduced phase transformation of PbTe nanocrystals with sizes of 13 and 5 nm up to ∼20 GPa was studied. Upon an increase of pressure, we observed that the 13 nm PbTe nanocrystals start a phase transformation from rocksalt structure to an intermediate orthorhombic structure and finally CsCl-type structure at 8 GPa, which is 2 GPa higher than that in bulk PbTe. In contrast, the 5 nm PbTe nanocrystals do not display the same type of transition with a further increased transition pressure as expected. Instead of orthorhombic or CsCl-type structure, the 5 nm PbTe nanocrystals turn to amorphous phase under a similar pressure (8 GPa). Upon a release of pressure, the 13 nm PbTe nanocrystals transform from high pressure CsCltype structure directly to rocksalt structure, whereas the 5 nm PbTe nanocrystals remain their amorphous phase to ambient conditions. The structure stability of rocksalt-type PbTe shows a significant reversal of Hall Petch effect. On the basis of such an observation with a critical size determination of ∼9 nm, PbTe nanocrystals appear as the first class of material that demonstrates a pressure-induced structural change from order to disorder. By sharing the insight of this reversed Hall Petch effect with associated transition types, we tuned our experimental protocol and successfully synthesized a sample with “high-pressure metastable structure”, amorphous phase at ambient pressure. This integrative study provides a feasible pathway to understand nucleation mechanism as a function of particle size and to explore novel materials with high-pressure metastable structure and unique properties under labaccessible conditions. KEYWORDS: Lead telluride, high pressure, phase transformation, pressure-induced amorphization, in situ X-ray diffraction, reversed Hall Petch effect

U

nderstanding of pressure-induced phase transformations and associated nucleation mechanisms of nanocrystals (NCs) has recently triggered significant interests not only in fundamental scientific research but also in practical applications in a variety of fields, from chemistry, physics, and materials science to geophysics.1 6 Previous studies on NCs under high pressures have revealed a series of interesting phenomena, including (1) the shift of the band gap energy in NCs to higher energy levels such as in PbSe NCs;7 (2) the onset of first-order phase transformation in NCs at elevated pressures such as in CdSe1 and CdS NCs;8 and (3) the enhancement of NC hardness and toughness such as in CdSe nanosheets stabilized with organic molecules.9 As subsequent applications, the adjustable band gap in semiconductor NCs enables development of pressure-tunable lasers or optical pressure sensing devices10 and increased structural stability as well as enhanced mechanical properties observed in NCs, technologically offering possibilities in the design and fabrication of hard and tough materials. Most recently, studies of pressure-induced orientation and attachment on Au11 and PbS12 NCs not only created a practical route in preparing one-dimensional (1D) or two-dimensional (2D) nanostructures but also provided insights of NC nucleation mechanisms under high pressure. r 2011 American Chemical Society

Additionally, pressure-induced amorphization (PIA), which is defined as a transformation of an ordered crystal to a disordered (or amorphous) phase upon compression is currently a subject of intense study. Although PIA has been observed in a wide spectrum of bulk materials,13 only a few studies on the basis of nanomaterials such as TiO214,15 and Y2O3 NCs16 were reported. The occurrence of PIA in these NCs displayed a strong sizedependence, that is, it can only be observed when the NC size is smaller than one certain value (designated as “critical size”). This “critical size” varies from one material to another, whose magnitude depends mostly on their crystal structure, chemical bonding, and phonon instability, etc. Determination of this effective critical size is extremely significant for not only understanding the dramatic changes on the chemical and physical properties in NCs but also being beneficial to fabrication of high pressure metastable materials that can be preserved at ambient conditions.16 It is accordingly essential to conduct a systematic investigation on the size-dependent PIA and structural stability to explore the “critical size”. Received: September 30, 2011 Revised: October 30, 2011 Published: November 01, 2011 5531

dx.doi.org/10.1021/nl203409s | Nano Lett. 2011, 11, 5531–5536

Nano Letters From an application point of view, the key purpose of highpressure study is aimed at discovering novel high-pressure metastable phases and their associated properties, as well as the nucleation mechanisms.9 It is well-known that manifestation of novel properties can be achieved by modification of either composition or structure. A high-pressure phase featuring a higher density of atoms compared to the starting material poses as a different structure. If a high-pressure metastable phase could be quenched and well preserved at ambient conditions, the manifested properties could be retained and technologically devised for potential and practical applications. Besides synthetic diamond from graphite, most high-pressure metastable phases observed so far can not be preserved after a pressure release. Experimental protocols in ambient environments necessary to prepare these materials with high pressure metastable structure have yet to be explored in depth by far. Lead telluride (PbTe) is a narrow direct band gap semiconductor with a large Bohr radius. These properties hold promise for many applications including thermoelectric, infrared detectors, and so forth.17 20 Moreover, PbTe has a relatively low phase transition pressure and temperature, making it an excellent candidate for possibly exhibiting both size- and pressure-tuned phenomena. Thus, a feasible model could be established toward guiding investigation of other materials. To the best of our knowledge, however, there has been no report of phase transformation study on PbTe NCs, although the phase stability and transformation of PbTe are important factors in practical applications such as building thermoelectric devices.21,22 Recent development of PbTe synthetic approaches has enabled a fabrication of highly monodisperse PbTe NCs with well-controlled morphology and particle size, providing access to systematically exploring size-dependent phase transformation and mechanisms under high pressure.18,23,24 In this work, we employed an in situ X-ray diffraction technique using a pair of diamond anvil cells (DAC) and investigated the pressure-induced phase transformation as a function of particle size in PbTe NC systems. Transmission electron microscopy (TEM) characterization provides additional assessment information of structure and texture in the recovered samples. Herein, we report our observations on size-dependent transition pressure and phase transformation, a reversal of Hall Petch effect on structural stability, as well as a PIA occurrence when the particle size is reduced to a critical value. On the basis of an understanding of the correlation between the transformation type and particle size, ultrasmall PbTe nanoparticles (NPs) were prepared using a slightly modified synthesis protocol and were shown to possess a high-pressure metastable phase. Experimental Section. Synthesis of PbTe NCs with Different Sizes. Chemicals. Diphenyl ether (DPE, 99%), oleic acid (OA, 90%), 1-octadecene (ODE, 90%), trioctylphosphine (TOP, 90%) and Te shots (99.999%) were obtained from Sigma-Aldrich, and lead oxide (PbO, 99.99%) was obtained from Alfa Aesar. Stock solutions of trioctylphosphine telluride (Te-TOP) with Te concentration of 1.0 and 0.5 M were prepared by mixing 0.1 and 0.05 mol of Te shots with 100 mL of TOP followed by stirring and heating at 150 °C overnight in a glovebox, respectively. Synthesis. The synthetic procedure of PbTe NCs with tunable particle sizes was adopted from previously established approaches.18,23,24 Synthesis strategy for ultrasmall NP preparation (3 nm) was further modified. All the experiments were carried out in a three-neck round-bottom flask equipped with a condenser and attached to a Schlenk line under an argon stream. The detailed recipes for all samples are given as follows.

LETTER

Scheme 1. Schematic of DAC and Experimental Configuration

13 nm PbTe NCs. 3 mmol of PbO, 6.0 mL of OA, and 14.0 mL of DPE were mixed in a flask, heated to 150 °C for 1 h and then cooled to room temperature, forming a Pb-oleate precursor solution. In another flask, after 3.0 mL of Te-TOP (1.0 M) and 14.0 mL of DPE were loaded and the mixture was heated and stabilized at 200 °C, the as-formed Pb-oleate solution was injected to make the reaction occurred. The system was then quenched 2 min later by promptly moving the flask into a cold water bath (special precaution should be taken). By adding a sufficient amount of anhydrous ethyl alcohol, the resultant PbTe NCs were isolated by centrifugation and then washed with hexane for two cycles. The purified NCs were finally dispersed in hexane, forming colloidal solutions. 5 nm PbTe NCs. 2 mmol of PbO, 3.8 mL of OA, and 12.0 mL of ODE were mixed in a flask, heated to 170 °C for 30 min, and then stabilized at 150 °C. Two milliliters of Te-TOP (0.5 M) was then injected into this solution, and the reaction was terminated 6 min later by rapidly replacing the heating mantle with a cold water bath. PbTe NCs were precipitated by adding methyl alcohol and acetone in sequence followed by centrifugation. Subsequent purification procedure is the same as above, and the NCs were eventually stored in hexane as colloidal suspensions. 3 nm PbTe NPs. 2 mmol of PbO, 1.4 mL of OA, and 40.0 mL of ODE were mixed and heated to 170 °C for 30 min in a flask, and the temperature was then stabilized at 140 °C. Two milliliters of Te-TOP (0.5 M) was subsequently injected into this mixture. Similarly, the reaction was terminated using a cold water bath 6 min later. Other post-treatment procedures were the same as those for the 5 nm NCs. In Situ High-Pressure Wide-Angle X-ray Diffraction (WAXRD) Experiments. A schematic diagram of DAC for in situ highpressure WAXRD experiments is illustrated in Scheme 1. Two gem-quality diamond anvils with each culet size of 500 μm were aligned and assembled into a DAC for pressurizing the samples, and a hole of 200 μm was drilled through a stainless Fe gasket and served as a sample chamber. PbTe NCs were directly deposited onto the tip of one diamond anvil by several cycles of drop-casting procedure, enabling a sufficient amount of PbTe NCs to accumulate in the sample chamber. After several ruby chips were placed on the top of the sample as pressure sensors, the other diamond anvil was accordingly used to cover the sample for pressurization. A standard ruby fluorescence technique was used to monitor the applied pressure on the sample.25 In situ high-pressure WAXRD experiments were conducted at room temperature using an angle dispersive synchrotron X-ray source at B2 station of Cornell High 5532

dx.doi.org/10.1021/nl203409s |Nano Lett. 2011, 11, 5531–5536

Nano Letters

LETTER

Figure 2. In situ WAXRD patterns of 5 nm PbTe NCs as a function of pressure (a), as well as 2D in situ WAXRD images of the 5 nm PbTe NCs under certain typical pressures (b d). Figure 1. In situ WAXRD patterns of 13 nm PbTe NCs during (a) compression and (b) decompression runs, respectively. The peaks marked with black triangles come from the diffraction of Fe gasket. (c) Graphic representation of the proposed structural transitions. Note: the molecular graphics were generated using the XCRYSDEN graphical package.27

Energy Synchrotron Source (CHESS).26 A double-bouncing Ge(111) monochromator converts incident white X-rays to a monochromatic beam at an optimized energy of 25.5 keV (equivalent to a wavelength of 0.48596 Å), which is parallel to the uniaxial compression direction (Scheme 1). A double pinhole tube cuts the X-ray beam down to 100 μm in diameter for illuminating the samples. The scattered signals from the sample are collected using a large area MAR345 detector, and then these original 2D images could be integrated into 1D X-ray diffraction patterns by means of a Fit2D software package. TEM Characterizations. The recovered sample from high pressure was carefully transferred from the gasket hole to a copper grid for TEM characterization. TEM images, high-resolution TEM (HRTEM) images, and selected-area electron diffraction (SAED) patterns were taken from a JEOL-2100F TEM instrument operated at 200 kV in the Small Scale Systems Integration and Packaging Center at State University of New York at Binghamton. Results and Discussion. Figure 1a,b shows several typical integrated WAXRD patterns of 13 nm PbTe NCs during compression and decompression runs collected under a pressure cycle of 0 20 GPa, respectively. It is obvious that the 13 nm PbTe NCs crystallize in a face-centered-cubic (fcc) structure (NaCltype or rocksalt) at ambient pressure (denoted as “0 GPa” in the corresponding patterns). The rocksalt structure of this sample remains stable up to a pressure of 8 GPa. However, all the X-ray diffraction peaks shift to higher 2θ angles upon compression, indicating a pressure-induced reduction of d-spacings or shrinkage of the unit cells. Above 8 GPa, the 13 nm PbTe NCs start a phase transformation from rocksalt structure to a mixture of intermediate orthorhombic (space group: Pnma) and CsCl-type structure.28 Both high pressure phases coexist until ∼14 GPa and then completely transform into a single CsCl-type structure under a maximum pressure of 20 GPa that was applied. Upon release of pressure,

the high pressure CsCl-type structure remains stable down to ∼1 GPa. Further reduction of pressure to ambient conditions results in an entire recovery of the original rocksalt structure without appearance of the intermediate orthorhombic phase. To clearly demonstrate this process, three structural polymorphs of the 13 nm PbTe NCs and their corresponding transition pressures are shown in Figure 1c. The pressure-induced sequence of structural change depicts an increase of atomic coordination number from six in rocksalt structure to seven in intermediate orthorhombic structure, and finally to eight in CsCl-type structure. Nevertheless, the translational symmetry is always being kept.13 An increase of atomic coordination tightens the crystal structure and enhances the atomic density, resulting in a significant improvement of hardness and strength. On the contrary, the reverse transformation from CsCl-type structure to rocksalt structure upon releasing pressure does not involve the intermediate phase, that is, 7-coordinated orthorhombic structure observed in the forward compression. More details about the asobserved phase transformation are presented using several typical 2D diffraction patterns provided in Figure S1 in the Supporting Information. In addition, the full width at halfmaximum (fwhm) for each corresponding peak in the WAXRD patterns of the starting and recovered samples appears almost unchanged (Figure S2 in Supporting Information), implying that a compression of the samples through various phase transformations did not cause a noticeable size change of particles. This observation suggests that the high pressure phase nucleation must take place in single domains. As is well-known, the crystal growth propagates much faster than a nucleation toward formation of a new phase. The single domain nucleation of high pressure phase in the 13 nm PbTe NCs does not produce multiple growth fronts. Therefore, neither approaching nor collision of propagating wave fronts breaks particles into small pieces. This event is different from the mechanism of multiple domain nucleation events that result in a breakdown of large particles in bulk materials due to the existence of multiple defects and dislocations. Previous high-pressure studies on bulk PbTe indicate that rocksalt phase starts the first-order transformation under 6 GPa,28 5533

dx.doi.org/10.1021/nl203409s |Nano Lett. 2011, 11, 5531–5536

Nano Letters

LETTER

Figure 4. Schematic of the relationship between the first-order transition pressure and particle size.

Figure 3. TEM images and HRTEM images of 5 nm PbTe NCs before (a and b in left panel) and after (c and d in right panel) high pressure treatment, respectively.

which is lower than the transition pressure of 8 GPa observed in this 13 nm PbTe NCs. Our result indicates that a reduction of particle size dramatically enhances the structure stability of PbTe NCs, which is consistent with the conclusion from other nanomaterials, such as CdSe,1 ZnS,6 TiO2,14,29 and so forth. This outcome implies that the size-tuned enhancement of structure stability follows the Hall Petch effect that has been observed over a wide range of NCs.30 Based on the Hall Petch effect, it is expected that a continuous decrease of particle size could further elevate transition pressure. Nevertheless, in situ WAXRD studies on 5 nm PbTe NCs provide a completely contrary result. Figure 2a presents several key WAXRD patterns of 5 nm PbTe NCs collected at the same pressure range (0 20 GPa). The starting sample possesses a rocksalt crystal structure, identical with the 13 nm PbTe NCs. However, its X-ray peaks are much broader than those of the 13 nm PbTe NCs (Figure S3 in Supporting Information, the disappearance of some minor X-ray peaks is mainly due to the reduction of particle size and resultant broadening-induced overlap of nearby peaks). WAXRD patterns of the 5 nm PbTe NCs (Figure 2a) show that the sample does not follow the Hall Petch prediction, that is, a continuous increase of the transition pressure over 8 GPa. Instead of this tendency, it remains stable in rocksalt structure to the same pressure of 8 GPa and then surprisingly starts a transformation to an amorphous phase rather than an orthorhombic or CsCl-type structure. A piece of most noticeable evidence is the appearance of a very broad diffuse scattering halo under pressures above 8 GPa (Figure 2b d), which is one typical feature for the onset of PIA. The disorder ratio increases progressively upon continuous pressurizing. When the pressure is greater than 12 GPa, only the broad peak located around 2θ = 8 11° remains, indicating that the starting NCs with rocksalt structure completely converts to an amorphous (disordered) phase. This amorphous phase was preserved upon release of pressure to ambient conditions (Figure 2a and

Figure S4 in Supporting Information). It can also be observed that the position of this broad peak around 2θ = 8 11° shifts to higher angles with further increasing pressure and then shifts to low angles after decompression. Such a phenomenon suggests that the local structure (e.g., Pb Te bonds) in the amorphous phase changes as a function of pressure. TEM observation provides additional information about the occurrence of PIA and the preservation of the as-formed amorphous phase at ambient conditions in the 5 nm PbTe NCs (Figure 3). The left panel of Figure 3 displays the TEM images of the 5 nm PbTe NCs before they were loaded to the DAC. Figure 3a reveals that these NCs are uniform with average size of 5 nm, and HRTEM image in Figure 3b shows their high crystallinity with groups of well-developed fringes, demonstrating a constant spacing of 0.34 nm that are ascribed to the (200) planes of rocksalt PbTe. The right panel of Figure 3 presents the features of the recovered sample from a high-pressure compression. Sheetlike nanostructure dominates the morphologies of the final products (Figure 3c). The HRTEM image in Figure 3d does not show any noticeable lattice fringe, confirming the occurrence of PIA, as well as the stability and preservation of amorphous structure at ambient condition. At the same pressure of 8 GPa, different types of phase transformations on the 13 and 5 nm PbTe NCs indicate a reversal of the Hall Petch effect on the structural stability of rocksalt PbTe. Although a reversal of the Hall Petch effect regarding hardness and strength has been observed in metallic NCs when particle size is smaller than a critical size of 10 nm,31,32 such a dramatic change on high pressure phase transformations was rarely reported. To understand these results, additional PbTe sample with an intermediate particle size (10 nm) was accordingly synthesized. The synthesis recipe of the 10 nm PbTe NCs is the same as that for the 13 nm NCs except a decrease of reaction temperature from 200 to 185 °C. The first-order transition pressure for this 10 nm sample was determined to be around 10 GPa, which is in accordance with the Hall Petch relation. In order to demonstrate the correlation between the first-order transition pressure and the particle size, a schematic based on these results as well as the reported transition pressure of 6 GPa from bulk PbTe28 is established in Figure 4. By applying Hall Petch relation, the critical size for PbTe NCs can be roughly determined as an intermediate size of around 9 nm (i.e., between 13 and 5 nm), which is very close to the critical size of the 10 nm observed from a wide range of NCs reasonably. It means that the first-order transition pressure increases with the 5534

dx.doi.org/10.1021/nl203409s |Nano Lett. 2011, 11, 5531–5536

Nano Letters

Figure 5. TEM image (a) and corresponding 2D WAXRD image (b) of ∼3 nm PbTe NPs before high pressure compression.

decrease of particle size and the maximum transition pressure could be allocated at the 9 nm PbTe NCs. If the size is continuously decreased below 9 nm, reversal of Hall Petch effect should occur, that is, smaller PbTe NCs should directly transform from a rocksalt structure to an amorphous form without showing orthorhombic phase. The smaller the particle size is, the lower the transition pressure is required. When particle size reduces to a specific value, this transition pressure should probably approach 0 GPa. In other words, when particle size is smaller than this value, an amorphous form should be preferred to nucleate and dominate the internal structure of NCs at ambient pressure, rather than a rocksalt structure. The reversal of the Hall Petch relation can be reasonably understood by a possible sliding and rotation of small NPs along the particle boundaries, similar to the mechanism given by Yip.32 Above the critical size, the blocking effect of particle boundaries on atomic movement dominates, accordingly enhancing the structural stability and mechanical performance; below the critical size, however, particles become much more active and the higher energy of the entire NP system can be reduced by a series of slight sliding, rotation, and ultimate fusion of NPs. The observed fusion of 5 nm CdTe NPs in such kind of study and the stress-driven formation of single crystal nanosheet from 3.5 nm PbS unequivocally support this proposed mechanism. To confirm this tendency, we accordingly modified our synthetic protocol and conducted several trial experiments (see details in Experimental Section). It turns out that we succeeded in synthesizing ultrasmall PbTe NPs with amorphous structure at ambient environments. A TEM image demonstrated in Figure 5a shows that the PbTe NPs have an average particle size of ∼3 nm. A typical 2D WAXRD image presented in Figure 5b demonstrates that the ∼3 nm PbTe NPs are of amorphous nature. The WAXRD patterns collected from the ∼3 nm PbTe NPs are similar to those of the pressure-induced amorphous phase either obtained under compression or quenched and preserved at ambient conditions (Figure S5 in Supporting Information). A reasonable agreement between our predictions and the practical synthesis allows us to confirm such a reversed Hall Petch effect. Thus, this work not only provides an understanding concerning the nature of both size- and pressure-tuned phenomena in PbTe NCs but also suggests an effective design for fabrication of novel and stable high-pressure structured materials at ambient conditions. We have systemically studied the size- and pressure-tuned phase transformation on PbTe NCs. Similar to bulk PbTe, the 13 nm PbTe NCs initially transform to an orthorhombic structure and then to a CsCl-type structure with a compression. The observed transition pressure (8 GPa), however, is significantly greater than the “bulk value” of 6 GPa. When the particle

LETTER

size reduces to 5 nm, a compression of PbTe NCs results in a direct formation to an amorphous phase under the same pressure of 8 GPa. These results suggest a significant reversed Hall Petch effect existing in small size rocksalt-type PbTe NCs. This is the first time that a dramatic change from an ordered to disordered structure under a compression was observed in small PbTe NCs. By following the discovery of the reversed Hall Petch effect, an ultrasmall PbTe NP synthesis protocol was developed and ∼3 nm PbTe NPs were produced. Such ultrasmall NPs were proven to possess a high-pressure metastable phase (amorphous phase) at ambient conditions. This study revealed a reversed Hall Petch effect in PbTe NCs and a critical size necessary to generate high-pressure metastable materials at ambient conditions, initializing a new research strategy consisting of an investigation among pressure, phase stability, and particle size effect.

’ ASSOCIATED CONTENT

bS

Supporting Information. Figures S1 S5. This material is available free of charge via the Internet at http://pubs.acs.org.

’ AUTHOR INFORMATION Corresponding Author

*E-mail: (J.F.) [email protected]; (Z.W.) [email protected].

’ ACKNOWLEDGMENT This project is supported by NSF (DMR-0731382). CHESS is supported by the NSF award DMR-0936384. The work at S3IP center at SUNY Binghamton was partially funded by New York State Office of Science Technology and Innovation, and the Empire State Development Corporation. We also thank Liang Li and Professor Guangwen Zhou from Department of Mechanical Engineering and Nathan Porter from Department of Chemistry, State University of New York at Binghamton for their contributions on construction of crystalline model and revision of the manuscript, respectively. ’ REFERENCES (1) Tolbert, S. H.; Alivisatos, A. P. Science 1994, 265, 373–376. (2) Guo, Q.; Zhao, Y.; Mao, W. L.; Wang, Z.; Xiong, Y.; Xia, Y. Nano Lett. 2008, 8, 972–975. (3) Zhang, H.; Gilbert, B.; Huang, F.; Banfield, J. F. Nature 2003, 424, 1025–1029. (4) Wang, Z. W.; Daemen, L. L.; Zhao, Y. S.; Zha, C. S.; Downs, R. T.; Wang, X. D.; Wang, Z. L.; Hemley, R. J. Nat. Mater. 2005, 4, 922–927. (5) Gilbert, B.; Huang, F.; Zhang, H.; Waychunas, G. A.; Banfield, J. F. Science 2004, 305, 651–654. (6) Srivastava, A.; Tyagi, N.; Sharma, U. S.; Singh, R. K. Mater. Chem. Phys. 2011, 125, 66–71. (7) Pietryga, J. M.; Zhuravlev, K. K.; Whitehead, M.; Klimov, V. I.; Schaller, R. D. Phys. Rev. Lett. 2008, 101, 217401. (8) Haase, M.; Alivisatos, A. P. J. Phys. Chem. 1992, 96, 6756–6762. (9) Wang, Z.; Wen, X. D.; Hoffmann, R.; Son, J. S.; Li, R.; Fang, C. C.; Smilgies, D. M.; Hyeon, T. Proc. Nat. Acad. Sci. U.S.A. 2010, 107, 17119–17124. (10) Podsiadlo, P.; Lee, B.; Prakapenka, V. B.; Krylova, G. V.; Schaller, R. D.; Demortiere, A.; Shevchenko, E. V. Nano Lett. 2010, 11, 579–588. (11) Wu, H.; Bai, F.; Sun, Z.; Haddad, R. E.; Boye, D. M.; Wang, Z.; Fan, H. Angew. Chem., Int. Ed. 2010, 49, 8431–8434. 5535

dx.doi.org/10.1021/nl203409s |Nano Lett. 2011, 11, 5531–5536

Nano Letters

LETTER

(12) Wang, Z.; Schliehe, C.; Wang, T.; Nagaoka, Y.; Cao, Y. C.; Bassett, W. A.; Wu, H.; Fan, H.; Weller, H. J. Am. Chem. Soc. 2011, 133, 14484–14487. (13) McMillan, P. F. J. Mater. Chem. 2004, 14, 1506–1512. (14) Swamy, V.; Kuznetsov, A.; Dubrovinsky, L. S.; McMillan, P. F.; Prakapenka, V. B.; Shen, G.; Muddle, B. C. Phys. Rev. Lett. 2006, 96, 135702. (15) Pischedda, V.; Hearne, G. R.; Dawe, A. M.; Lowther, J. E. Phys. Rev. Lett. 2006, 96, 035509. (16) Wang, L.; Yang, W.; Ding, Y.; Ren, Y.; Xiao, S.; Liu, B.; Sinogeikin, S. V.; Meng, Y.; Gosztola, D. J.; Shen, G.; Hemley, R. J.; Mao, W. L.; Mao, H. K. Phys. Rev. Lett. 2010, 105, 095701. (17) Shchennikov, V. V.; Ovsyannikov, S. V. Solid State Commun. 2003, 126, 373–378. (18) Lu, W.; Fang, J.; Stokes, K. L.; Lin, J. J. Am. Chem. Soc. 2004, 126, 11798–11799. (19) Quan, Z.; Valentin-Bromberg, L.; Loc, W. S.; Fang, J. Chem. Asian J. 2011, 6, 1126–1136. (20) Wise, F. W. Acc. Chem. Res. 2000, 33, 773–780. (21) Wang, Y.; Chen, X.; Cui, T.; Niu, Y.; Wang, Y.; Wang, M.; Ma, Y.; Zou, G. Phys. Rev. B 2007, 76, 155127. (22) Xu, L.; Zheng, Y.; Zheng, J.-C. Phys. Rev. B 2010, 82, 195102. (23) Zhang, J.; Kumbhar, A.; He, J.; Das, N. C.; Yang, K.; Wang, J.-Q.; Wang, H.; Stokes, K. L.; Fang, J. J. Am. Chem. Soc. 2008, 130, 15203–15209. (24) Murphy, J. E.; Beard, M. C.; Norman, A. G.; Ahrenkiel, S. P.; Johnson, J. C.; Yu, P.; Micic, O. I.; Ellingson, R. J.; Nozik, A. J. J. Am. Chem. Soc. 2006, 128, 3241–3247. (25) Mao, H. K.; Xu, J.; Bell, P. M. J. Geophys. Res. 1986, 91, 4673–4676. (26) Wang, Z. W.; Chen, O.; Cao, C. Y.; Finkelstein, K.; Smilgies, D. M.; Lu, X. M.; Bassett, W. A. Rev. Sci. Instrum. 2010, 81, 093902. (27) Kokalj, A. J. Mol. Graphics Modell. 1999, 17, 176–179. (28) Rousse, G.; Klotz, S.; Saitta, A. M.; Rodriguez-Carvajal, J.; McMahon, M. I.; Couzinet, B.; Mezouar, M. Phys. Rev. B 2005, 71, 224116. (29) Wang, Z.; Saxena, S. K. Solid State Commun. 2001, 118, 75–78. (30) Wang, Z.; Guo, Q. J. Phys. Chem. C 2009, 113, 4286–4295. (31) Yip, S. Nat. Mater. 2004, 3, 11–12. (32) Yip, S. Nature 1998, 391, 532–533.

5536

dx.doi.org/10.1021/nl203409s |Nano Lett. 2011, 11, 5531–5536