Chatelaine, Aug 2011

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

NANO LETTERS

Insulator-to-Metal Transition in Nanocrystal Assemblies Driven by in Situ Mild Thermal Annealing


Paul Beecher, Aidan J. Quinn, Elena V. Shevchenko, Horst Weller, and Gareth Redmond*,
Nanotechnology Group, NMRC, Lee Maltings, Prospect Row, Cork, Ireland, and Institute of Physical Chemistry, UniVersity of Hamburg, D-20146 Hamburg, Germany
Received April 30, 2004; Revised Manuscript Received June 3, 2004

2004 Vol. 4, No. 7 1289-1293

ABSTRACT
We report on exploration of internanocrystal coupling in assemblies of CoPt3 nanocrystals (i.e., artificial atom solids) probed by variable temperature charge transport measurements. Nanocrystal devices can be tuned in situ from Mott insulating to metallic behavior using mild thermal annealing. Thermal analysis suggests that the observed insulator-to-metal transition is due to melting and reorganization of the nanocrystal ligand shells causing a reduction in mean internanocrystal distance and is not due to nanocrystal ligand desorption or nanocrystal sintering.

Metal and semiconductor nanocrystals with diameters in the range 1-30 nm are the subject of intense research interest since these materials can behave as artificial atoms or quantum dots, with size-dependent optical and electronic properties bridging those of bulk and molecular materials, e.g., plasmon resonances, single charge tunneling, and quantum confinement.1 The unique self-assembly properties of organically passivated monodisperse nanocrystals allow two- and three-dimensional (2D, 3D) nanocrystal arrays to be formed. Such arrays can exhibit collective behavior if individual nanocrystals couple to neighbors and therefore offer the promise of artificial atom solids with novel electronic and/or optical properties derived from both the size-dependent electronic properties of the individual nanocrystals and the internanocrystal electronic coupling mechanisms. Understanding these coupling mechanisms and their influence on the collective behavior of nanocrystal assemblies is of critical importance if nanocrystal-based architectures with tunable optical and electronic properties are to become a reality. Internanocrystal electronic coupling in metal and semiconductor nanocrystal assemblies has previously been explored by optical absorption and energy transfer,2 secondharmonic generation,3 photoconductivity,4,5 and charge transport measurements.7-13 Effects resulting from tuning the internanocrystal separation include observations of transitions from localized to collective behavior in assemblies of semiconductor nanocrystals2 and demonstrations of insulator* Corresponding author. Email: gareth.redmond@nmrc.ie Nanotechnology Group, NMRC. University of Hamburg. 10.1021/nl049358+ CCC: $27.50 Published on Web 06/22/2004 2004 American Chemical Society

to-metal transitions in compressed 2D Langmuir-Blodgett monolayers of silver nanocrystals at room temperature as the internanocrystal separation was reduced.3,6 In this work, we investigate internanocrystal coupling by measuring the charge transport characteristics of nanocrystal arrays self-assembled between electrodes (interelectrode spacings 70 nm) as a function of the post-deposition anneal temperature. We show that mild thermal processing can be exploited to tune internanocrystal coupling in these laterally contacted nanocrystal assemblies progressively from Mott insulating through to metallic regimes without nanocrystal ligand desorption or sintering of individual nanocrystals. This represents a key milestone in the development of functional nanocrystal-based architectures. CoPt3 nanocrystals (3.8 nm in diamater) stabilized by 1-adamantanecarboxylic acid and hexadecylamine ligands have been synthesized as reported previously,14 via a modified polyol process in a high boiling point coordinating solvent.15 As-synthesized nanocrystals oxidize very slowly (air-stable for periods of months), are monodisperse ( ) 9% without post-synthesis size-selection) and highly crystalline, and possess a chemically disordered fcc structure.14 Excess organic stabilizers are removed by purification cycles of nonsolvent (ethanol) addition to the nanocrystal solution, centrifugation of the precipitate, and redispersion of the resultant nanocrystal material in toluene. Owing to their narrow size distribution, multilayer nanocrystal arrays deposited on carbon-coated copper TEM grids pack into ccplike superlattices exhibiting long-range order when nanocrystal solutions are dropped onto substrates and the toluene

Figure 1. (a) TEM image of an array of as-synthesized 3.8 nm diameter CoPt3 nanocrystals on a carbon-coated copper grid. Inset: Fourier transform indicating multiple domain formation. (b) High-resolution SEM of a typical nanocrystal device assembled between gold electrodes on an oxidized silicon substrate.

solvent is allowed to evaporate in a controlled manner; see Figure 1a (CM-300 TEM, Philips Electronics N.V). The nanocrystal cores are highly crystalline and are separated by the hexadecylamine and bulky 1-adamantanecarboxylic acid ligands coordinated to each nanocrystal surface.14 Values for the mean internanocrystal separation, s, can be calculated by generating numerical Fourier transforms of TEM images from regions typically containing several hundred nanocrystals; see Figure 1a inset. The Fourier data typically show ring-like patterns of dots, suggesting formation of multiple domains within each region. From Fourier analysis of TEM data for several regions, we calculate s 2.3 nm. More crude estimates for s can also be extracted from TEM (or SEM) data by counting the number of nanocrystals in a defined area and calculating the mean area occupied by a single nanocrystal, ANC.5 From analysis of several TEM images for regions similar to that shown in Figure 1a (which has 135 nanocrystals in a 4700 nm2 region), we estimate ANC ) 35 ( 2 nm2 for arrays of 3.8 nm diameter nanocrystals deposited on TEM grids. Assuming hexagonal packing, ANC ) ( 3/2)(d+s)2, yielding s 2.5 nm, in reasonable agreement with values calculated from Fourier transform data. For measurements of electrical transport through nanocrystal assemblies, interdigitated finger electrodes 4 m in
1290

length with interelectrode gaps 70 nm are fabricated on thermally oxidized silicon wafer chips using electron beam lithography followed by metal evaporation (Cr 5 nm/Au 25 nm) and liftoff. Multilayer nanocrystal assemblies are deposited between the electrodes by dropping 5 L of a purified nanocrystal solution onto each electrode array under ambient conditions and allowing the solvent to evaporate slowly. Figure 1b shows a high resolution SEM of a typical as-deposited nanocrystal device (JSM-6700F, JEOL UK Ltd.). The nanocrystal assembly has a low void density (1%) but is not well ordered, presumably as a result of structural perturbations due to the electrodes. From analysis of multiple SEM images of nanocrystal assemblies deposited between electrodes on oxidized silicon substrates, we estimate ANC ) 36 (2 nm2 (70 nanocrystals in regions of area 2500 nm2) similar to the values extracted from TEM data for carbon substrates (ANC ) 35 ( 2 nm2). Electrical measurements of nanocrystal devices are performed using a semiconductor parameter analyzer (HP4156A, Agilent Technologies, U.K.), with voltage resolution <100 V and current resolution <100 fA. For room-temperature measurements, electrode chips are contacted using a probe station (PML-8000, Wentworth Labs, U.K.). Measured roomtemperature low-bias resistances, R, of as-deposited devices always exceed 1 T for multilayer nanocrystal arrays selfassembled between electrodes (70 nm gap), consistent with large internanocrystal separations. A range of in situ thermal anneals has been developed to increase the internanocrystal coupling through partial melting and reorganization of the protecting ligands.12,16 Post-deposition annealing is performed under reducing conditions (5% H2, 95% N2 for 1 h) to reduce any surface oxide layer which may form during handling. In this regard, previous XRD, TEM, and solubility studies have shown that these 3.8 nm diameter CoPt3 nanocrystals do not show evidence of significant oxidation following anneals at temperatures up to 200 C in air over 2 h,14 in contrast to surfaces of pure Co colloids which rapidly develop surface oxide layers up to several monolayers thick under ambient conditions.17 The maximum anneal temperature is deliberately limited to 150 C to minimize ligand desorption and to prevent nanocrystal sintering.18 In contrast to other reports, we aim to deliberately avoid altering the structure and chemistry of the individual nanocrystals and protecting ligands, so that we can be confident that changes in the charge transport characteristics are solely due to increased internanocrystal electronic coupling.5,10,15 The nanocrystal array anneal temperature, Ta, has a profound effect on the measured device electrical characteristics. For devices annealed at Ta ) 80 C, post-anneal I-V characteristics measured at room temperature exhibit weak nonlinearity with a low bias resistance, R 100 M, over 4 orders of magnitude lower than values measured for unannealed devices; see Figure 2a. Taking the assembly dimensions as 4000 70 30 nm3, i.e., assuming the arrays completely fill the interelectrode gap, yields an order-ofmagnitude estimate for the device resistivity, F 102 m. For variable temperature measurements (3 - 400 K), electrode chips are mounted in leadless chip carriers, wireNano Lett., Vol. 4, No. 7, 2004

Figure 2. (a) Post-anneal variable temperature current-voltage characteristics for a device annealed at Ta ) 80 C. Inset: Sequential I-V curves measured at 3 K demonstrating reproducible voltage threshold for current suppression, VT ) 0.24 V. (b) Semilog plot of device resistance, R vs 1/T. Data (circles) are well fit by activation model (solid line) for kBT < Ec (Ec ) 30 meV). (c) Measured low temperature (3.1 K) scaling behavior of current above Coulomb blockade threshold voltage, VT, for three devices. The values of the extracted scaling exponents, , suggest conduction through current-carrying networks with dimensionality g2D.

bonded and inserted into a liquid helium bath cryostat (Spectrostat, Oxford Instruments, U.K.). The measured low-bias resistance, R, increases monotonically with decreasing temperature, T, and provides strong evidence for the size-monodispersity of the nanocrystal array and the stability of the protecting ligand monolayer following the anneal. The data are consistent with tunnel transport through an array of size-similar nanoscale metallic islands in an insulating matrix, where the carrier density is activated, with activation energy Ec.19 Figure 2a (inset) presents a semilog plot of R vs 1/T for the device with I-V data shown in Figure 2a, demonstrating that the device exhibits activated behavior in the temperature range 20-80 K, R exp[Ec/kBT]) for kBT < Ec where kB is the Boltzmann constant.7,9,10,12 For all such measured devices, data fits yield activation energies in the range 28 < Ec < 34 meV. The narrow range of measured activation energies also provides further supporting evidence of the monodispersity of the nanocrystals, previously measured by TEM.14 The total capacitance of a nanocrystal in the assembly due to its nearest neighbors, Cnn, is given by Ec ) e2/2Cnn. For a 2-3D film with ccp packing, each nanocrystal will have
Nano Lett., Vol. 4, No. 7, 2004

between 6 and 12 nearest neighbors. Taking an activation energy for nanocrystal charging Ec ) 31 meV, the total capacitance of each nanocrystal is Cnn 2.6 aF, and the pairwise internanocrystal capacitance is estimated to be C ) Cnn/9 0.3 aF. The mean post-anneal internanocrystal spacing, sPA*, for the device with I/V data shown in Figure 2a may be calculated from the pairwise internanocrystal capacitance using C 2 0 r ln(1+2r/sPA*), where r is the nanocrystal radius, r ) 1.9 nm, and is the relative permittivity of the ligand, ( 2.7).10,20,21 A value of sPA* 2.2 nm is estimated based on the measured activation energy for nanocrystal charging Ec ) 31 meV. Therefore, both the magnitude and temperature dependence of the postanneal measured low-bias resistance for devices annealed at Ta ) 80 C suggest that the mean internanocrystal separation has decreased. Concerning the voltage dependence of the current below 35 K for the device with I-V data shown in Figure 2a, complete current suppression is observed for voltages below the threshold voltage, VT ) 0.24 V. Measured values of VT are device dependent, 0.24 V < VT < 0.8 V, probably due to local geometric disorder and/or charge disorder arising from random parasitic charges present within the nanocrystal arrays and/or the oxidized silion substrates. For voltages above VT, theory predicts that for a uniform array of identical nanoscale metallic islands separated by tunnel barriers, the current should follow a power law, I (V/VT -1), where the scaling exponent, , depends on the dimensionality of the array, i.e., the number of accessible current paths.10,11,22 Theory predicts ) 5/3 for 2D networks while numerical simulations have yielded values of 2 ( 0.2.22 Figure 2b presents log-log plots of data measured for three CoPt3 nanocrystal devices which demonstrate that the current through the devices follows a power law over 3 orders of magnitude with calculated values of in the range 2.2 < < 2.8 (/ 12%). The magnitude and variation of the value of the scaling exponent for the devices is indicative of transport through networks of conduction paths in the different CoPt3 nanocrystal arrays with dimensionality equal to or greater than 2D. These results are consistent with the nanocrystal multilayer array deposition conditions employed during device fabrication and SEM images of typical devices; see Figure 1b. Overall, the results agree with previous experimental studies on monodisperse nanocrystal arrays self-assembled between nanoscale electrodes, where values of 2.2 < < 2.7 have been reported for multilayer arrays of 10 nm Co nanocrystals and ) 2.8 for multilayer arrays of 10.2 nm CoPt3 nanocrystals.10,12 We conclude that 3.8 nm nanocrystal array devices annealed at Ta ) 80 C are Mott insulators with charging energies governed by the nanocrystal size and also internanocrystal capacitances arising from reduced internanocrystal separation. Unannealed device arrays are also expected to behave as Mott insulators. However, values for the nanocrystal charging energy in such arrays cannot be determined due to the extremely high (>T) device resistance arising from larger mean internanocrystal separations measured for these arrays. Increasing the anneal temperature to Ta ) 100 C has a pronounced effect on the measured transport properties of
1291

Figure 3. (a) Measured low-bias device resistance, R vs T, for a device annealed at Ta ) 100 C, demonstrating intrinsic metallic behavior with mild disorder. Inset: Measured I-V curves are linear across the entire bias voltage range ((1 V) at all measured temperatures, indicating enhanced internanocrystal coupling relative to devices annealed at Ta ) 80 C. (b) Measured R vs T for a device annealed at Ta ) 150 C, showing metallic behavior at all measurement temperatures. Inset: Measured I-V curves are linear at all measured temperatures. The bias voltage is limited to (0.1 V to minimize Joule heating effects.

nanocrystal array devices. The post-anneal I-V characteristics are linear over the entire measured voltage range ((1 V) at all measured temperatures, in contrast to the nonlinearity observed over the same voltage range for devices annealed at Ta ) 80 C; see Figure 3a (inset). This suggests that activated tunneling is no longer the dominant transport process for these devices. Moreover, the low-bias resistance measured at room temperature, RRT 190 k (corresponding resistivity F 10-1 m), is 3 orders of magnitude lower than values measured for devices annealed at Ta ) 80 C. Both the magnitude of the room-temperature device resistance and the linearity of device I-V characteristics at all measurement temperatures suggest a strong enhancement of the internanocrystal coupling. We propose that this is due to a further anneal-induced reduction in the internanocrystal separation. In terms of a Mott-Hubbard framework, at sufficiently large internanocrystal separations where nanocrystal charging dominates over exchange coupling, e.g., for unannealed devices or devices annealed at 80 C, the nanocrystal array devices behave as Mott insulators where the Mott-Hubbard gap UG is on the order of the nanocrystal charging energy, i.e., UG g 30 meV. As the internanocrystal separation is reduced, the Hubbard gap narrows due to
1292

increased exchange coupling and finally closes at the insulator-metal transition.23 Variable temperature measurements of R vs T for Ta ) 100 C show a small positive temperature coefficient of resistance (TCR) from 300 - 80 K, characteristic of weakly metallic behavior; see Figure 3a. Within the Mott-Hubbard framework, the positive TCR data suggest that the mean internanocrystal separation within the device array has been reduced following the 100 C anneal and that, as a consequence, the Hubbard gap is closed. A minimum measured resistance is observed at Tmin ) 80 K and a small negatiVe TCR is observed from 80 - 4 K, suggesting the presence of geometric disorder within the nanocrystal array. However, the negative TCR data cannot be fit using activation, hopping, or power-law models.19,23,24 In fact, the variation in measured resistance vs temperature, R(T) ) |R(T) - R(Tmin)|/R(Tmin), is less than 7% across the entire temperature range, suggesting that the disorder, though finite, is sufficiently small and the localization lengths are sufficiently large that the nanocrystal array is on the metallic side of the metal-insulator boundary. Negative-positive TCR transitions have been observed previously in micronscale arrays of multilayers and compressed Langmuir monolayers of 7 nm diameter Ag nanocrystals, as well as in highly doped polymers.8,9,24 For these reports, the transitions were observed at higher temperatures (Tmin > 200 K) and the relative increases in measured resistance vs temperature below Tmin, R(T), were several orders of magnitude larger than in this work, indicating substantial contributions from disorder-induced localization. Increasing the anneal temperature to Ta ) 150 C produces devices with measured room-temperature resistances 2 orders of magnitude lower than for Ta ) 100 C, R 1 k (corresponding resistivity F 10-3 m). Variable temperature electrical characterization of devices annealed at Ta ) 150 C show linear I-V curves over a wide temperature range (4 K < T < 300 K); see Figure 3b (inset). The low bias resistances, R, monotonically increase with temperature, indicating metallic behavior at all temperatures; see Figure 3b. The estimated resistivity value is comparable to reported values for assemblies of 4 nm gold nanocrystals linked by 1.1 nm long bis-dithiocarbamate deriviatives,13 and is over 3 orders of magnitude larger than measured values for ultrathin amorphous Pt films (of thickness 0.6 nm).25 The data indicate that our devices show greatly enhanced electron scattering relative to continuous nanoscale metal films, as expected for an array of strongly coupled (but distinct) metal islands. Thermal data support our assertion that the nanocrystal cores and protecting ligand monolayers are not compromised by the mild annealing temperatures employed in our studies and that the dramatic reduction in device resistance with anneal temperature is the result of increased internanocrystal coupling, rather than formation of local high-conductance paths through the array due to ligand desorption and/or nanocrystal sintering. Figure 4 shows simultaneous thermogravimetric (TGA) and differential thermal analysis (TGADTA) studies on mg quantities of 4 nm diameter CoPt3 nanocrystals synthesized and purified using the same procedures
Nano Lett., Vol. 4, No. 7, 2004

Acknowledgment. The authors thank John Rea for technical assistance, Gianluca de Marzi for SEM imaging, Liam Floyd for TEM image analysis, Sylvia Williamson of the University of St. Andrews, U.K. for TGA-DTA measurements, and Oliver Harnack of SONY Europe for assistance in fabrication of the electrodes. This work was supported by the EU under the BIOAND project (IST-1999-11974) and by the Irish HEA PRTLI-1 Nanoscale Science and Technology Initiative. References
(1) Collier, C. P.; Vossmeyer, T.; Heath, J. R. Annu. ReV. Phys. Chem. 1998 49, 371. Murray, C. B.; Kagan, C. R.; Bawendi, M. G. Annu. ReV. Mater. Sci. 2000 30, 545. (2) Kagan, C. R.; Murray, C. B.; Nirmal, M.; Bawendi, M. G. Phys. ReV. Lett. 1996, 76, 1517. Artemyev, M. V.; Bibik, A. I.; Gurinovich, L. I.; Gaponenko, S. V. Woggon, U. Phys. ReV. B 1999, 60, 1504. Artemyev, M. V.; Woggon, U.; Jaschinski, H.; Gurinovich, L. I.; Gaponenko, S. V. J. Phys. Chem. B 2000, 104, 11617. Micic, O. I.; Ahrenkiel, S. P.; Nozik, A. J. Appl. Phys. Lett. 2001, 78, 4022. (3) Collier, C. P.; Saykally, R. J.; Shiang, J. J.; Henrichs, S. E.; Heath, J. R. Science 1997, 277, 1978. (4) Leatherdale, C. A.; Kagan, C. R.; Morgan, N. Y.; Empedocles, S. A.; Kastner, M. A.; Bawendi, M. G. Phys. ReV. B 2000, 62, 26699. (5) Drndic, M.; Jarosz, M. V.; Morgan, N. Y.; Kastner, M. A.; Bawendi, M. G. J. Appl. Phys. 2002, 92, 7498. (6) Markovich, G. Collier, C. P.; Heath, J. R. Phys. ReV. Lett. 1998, 80, 3807. (7) Andres, R. P.; Bielefeld, J. D.; Henderson, J. I.; Janes, D. B.; Kolagunta, V. R.; Kubiak C. P.; Mahoney, W.; Osifchin R. G. Science 1996, 273, 1690. Brust M.; Bethell D.; Kiely C. J.; Schiffrin D. J. Langmuir 1998, 14, 5425. Wuelfing W. P.; Green S. J.; Pietron J. J.; Cliffel D. E.; Murray R. W. J. Am. Chem. Soc. 2000, 122, 11465. (8) Doty, R. C.; Yu, H.; Shih, C. K.; Korgel, B. A. J. Phys. Chem. B 2001, 105, 8291. (9) Sampaio, J. F.; Beverly, K. C.; Heath, J. R. J. Phys. Chem. B 2001, 105, 8797. Beverly, K. C.; Sampaio, J. F.; Heath, J. R. J. Phys. Chem. B 2002, 106, 2131. (10) Black, C. T.; Murray, C. B.; Sandstrom, R. L.; Sun, S. Science 2000, 290, 1131. (11) Parthasarathy, R.; Lin, X.-M.; Jaeger, H. M. Phys. ReV. Lett. 2001, 87, 186807. (12) Beecher, P.; Quinn, A. J.; Shevchenko, E. V.; Weller, H.; Redmond, G. J. Phys. Chem. B 2004, in press. (13) Wessels, J. M.; Nothofer, H.-G.; Ford, W. E.; von Wrochem, F.; Scholz, F.; Vossmeyer, T.; Schroedter, A.; Weller, H.; Yasuda, A. J. Am. Chem. Soc. 2004, 126, 3349. (14) Shevchenko, E. V.; Talapin, D. V.; Rogach, A. L.; Kornowski, A.; Haase, M.; Weller, H. J. Am. Chem. Soc. 2002, 124, 11480. (15) Sun, S.; Murray, C. B.; Weller, D.; Folks, L.; Moser, A. Science 2000, 287, 1989. (16) Meulenberg, R. W.; Strouse, G. F. J. Phys. Chem. B 2001, 105, 7438. Sandhyarani, N.; Pradeep, T.; Chakrabarti, J.; Yousuf, M.; Sahu, H. K. Phys. ReV. B 2000, 62, R739. Mitra, S.; Nair, B.; Pradeep, T.; Goyal, P. S.; Mukhopadhyay, R. J. Phys. Chem. B 2002, 106, 3690. Korgel, B. A. Phys. ReV. Lett. 2001, 86, 127. (17) Wiedwald, U.; Spasova, M.; Farle, M.; Hilgendorff, M.; Giersig, M. J. Vac. Sci. Technol. A 2001, 19, 1773. Sobal, N. S.; Hilgendorff, M.; Mohwald, H.; Giersig, M.; Spasova, M.; Radetic, T.; Farle, M. Nano Lett. 2002, 2, 621. (18) Martin, J. E.; Odinek, J.; Wilcoxon, J. P.; Anderson, R. A.; Provencio, P. J. Phys. Chem. B 2003, 107, 430. (19) Neugebauer, C. A.; Webb, M. B. J. Appl. Phys. 1962, 33, 74. (20) Rampi, M. A.; Schueller, O. J. A.; Whitesides, G. M. Appl. Phys. Lett. 1998, 72, 1781. (21) Laiktman, B.; Wolf, E. L. Phys. Lett. A 1989, 139, 257. (22) Middleton, A. A.; Wingreen, N. S. Phys. ReV. Lett. 1993, 71, 3198. (23) Mott, N. F. Metal-Insulator Transitions, 2nd ed.; Clarendon Press: Oxford, 1997. (24) Reghu, M.; Cao, Y.; Moses, D.; Heeger, A. J. Phys. ReV. B 1993, 47, 1758. (25) Fischer, G.; Hoffmann, H.; Vancea, J. Phys. ReV. B 1980, 22, 6065.

Figure 4. Simultaneous thermogravimetric and differential thermal analysis (TGA, DTA) performed on 4 nm CoPt3 nanocrystals. Substantial ligand desorption and nanocrystal sintering occur only at temperatures well in excess of the maximum device anneal temperature employed in this work (150 C).

employed for the 3.8 nm diameter nanocrystals used in our devices (SDT 2960, TA Instruments, U.K.). No substantial mass reduction is observed in TGA data at temperatures below 250 C, indicating that substantial ligand desorption occurs only above 250 C. Further, the DTA data demonstrate that nanocrystal core melting occurs only at temperatures in excess of 300 C and appears to be driven by ligand desorption. Previous XRD and solubility data also confirm the stability of the nanocrystals in air for anneal temperatures up to 200 C.14 Therefore, the net effect of premeasurement thermal annealing of devices is a reduction in internanocrystal spacing through partial melting and reorganization of the protecting ligands followed by freezing of the nanocrystal solid upon cooling to room temperature.12,16,18 We conclude that anneals at Ta ) 80 C yield devices with internanocrystal separations small enough to allow measurable tunneling but large enough so that the charging energy is much greater than the internanocrystal exchange coupling, thus these devices are Mott insulators with a Coulomb gap on the order of the charging energy. The activated resistance vs temperature data implies a sizeuniform nanocrystal array and the scaling dependence of the low-temperature I-V data suggests 2D/3D networks of current-carrying paths. Raising the anneal temperature to Ta ) 100 C enhances the internanocrystal coupling to a degree where the charging energy no longer dominates and exchange coupling is sufficiently strong to induce disordered metallic behavior, i.e., formation of delocalized (band-like) states. In terms of a Mott-Hubbard band picture, this corresponds to a band-crossing insulator-metal transition where devices annealed at Ta ) 80 C are Mott insulators, annealing at Ta ) 100 C closes the Hubbard gap (with a small but finite density of disordered states at the band-crossing point), and annealing at Ta ) 150 C produces coupling strong enough to delocalize electrons across the nanocrystal array at all temperatures resulting in metallic behavior. We believe this constitutes the first report of an in situ anneal-induced insulator-to-metal transition in metal nanocrystal arrays.
Nano Lett., Vol. 4, No. 7, 2004

NL049358+
1293

You might also like