Surface Modification of Zinc Oxide Nanoparticles Influences the

Jun 23, 2007 - (a) Beck, A.; Bednorz, J. G.; Gerber, C.; Rossel, C.; Widmer, D. Appl. Phys. Lett. 2000, 77, 139. [Crossref], [CAS]. (2) . Reproducible...
0 downloads 0 Views 196KB Size
10150

2007, 111, 10150-10153 Published on Web 06/23/2007

Surface Modification of Zinc Oxide Nanoparticles Influences the Electronic Memory Effects in ZnO-Polystyrene Diodes Frank Verbakel, Stefan C. J. Meskers,* and Rene´ A. J. Janssen Molecular Materials and Nanosystems, EindhoVen UniVersity of Technology, P.O. Box 513, 5600 MB EindhoVen, The Netherlands, and Dutch Polymer Institute, P.O. Box 902, 5600 AX EindhoVen, The Netherlands ReceiVed: April 18, 2007; In Final Form: May 25, 2007

Resistive switching effects in diode structures with a spin-coated active layer containing zinc oxide (ZnO) nanoparticles in a polystyrene matrix are studied. The switching effect can be influenced by modification of the surface of the nanoparticles with coordinating ligands (amines and thiols). Using n-propylamine as a ligand, memory effects are observed without the diodes having undergone the forming step that is usually required before switching effects can be observed in bulk metal oxides. Memory effects are characterized by impedance spectroscopy and temperature-dependent current-voltage measurements and involve a spontaneous, thermally activated gradual transition from a state with high frequency independent conduction to a state with lower conductivity.

Introduction Resistive switching in metal oxides embedded in metalinsulator-metal (MIM) structures is well-known1 and is currently intensely investigated, partly because of its potential application in data storage.2 Research has focused on switching in bulk oxide layers that are either amorphous, microcrystalline, porous, and nonstoichiometric1b,c or inhomogeneous.3 Different mechanisms have been proposed to explain the resistive switching phenomena.4 A large number of reports describe electronic transport via filaments that arise during a forming step of the MIM diode.1,5 Because of the structural disorder and spontaneous formation of filamentary conduction paths, the switching process is difficult to control, as it is influenced by random variations in the insulating layer. In this paper, we show that the resistive switching effect in nanostructured oxide layers, made by spin-coating oxide nanoparticles from solution, can be influenced via surface chemistry of the nanoparticles. This shows that knowledge from colloid and surface science of oxide nanoparticles may be used in the design of materials for the next generation memory technology.6 We have reported on memory effects in MIM structures with an insulating layer consisting of ZnO nanoparticles mixed with polystyrene (PS).7 The use of two different electrodes (aluminum and poly(3,4-ethylenedioxy thiophene):polystyrene sulfonate (PEDOT:PSS)) results in diode-like current voltage characteristics for the pristine device. After a forming step induced by the application of forward bias voltage, the resistance of these devices drops considerably and can be reversibly altered by applying voltage pulses of different polarity. These switching effects resemble those reported for bulk ZnO,8,9 but the mechanism responsible for switching is not yet known for either the bulk or the nanostructured solid. The use of oxide nanoparticles offers the interesting possibility for chemical modifica* Corresponding author. [email protected].

10.1021/jp072999j CCC: $37.00

tion of the surface of the ZnO particle. Surface modification of ZnO nanoparticles is known, and binding of thiols,10 acids,11 and amines12 has been reported. The surface of the particles is expected to play an important role in determining the electrical properties of the blend because conduction of charge occurs through interparticle contacts.13 Experimental Section The diodes consist of a spin-coated active layer between two electrodes (Figure 1). The bottom electrode is tin-doped indium oxide (ITO) covered with a thin spin-coated film of PEDOT: PSS of 0.6 × 102 nm. The ZnO nanoparticles were synthesized by reacting zinc acetate with potassium hydroxide in methanol.14 ZnO nanoparticles (np) with an average diameter of 5 nm are precipitated, washed, centrifuged, and finally dissolved in chloroform. The active layer is spin-coated at 1500 rpm from a solution containing 20 mg PS (Mw ) 250 kg/mol, PDI ) 2.2), 10 mg of ZnO np and ligand per milliliter of chloroform. For n-propylamine (PA), the ratio is PS:ZnO:PA ) 4:2:1 (by weight). The other ligands were added to the solution in such an amount that the molar concentration of the ligand was the same as for PA. The addition of ligand does not influence the film thickness (4 × 102 nm). The top electrode is made by evaporation of aluminum (1 × 102 nm). The active device area is 9.5 mm2. After deposition of Al, the devices are stored and characterized in an inert atmosphere (O2, H2O e 1 ppm) or in vacuum to prevent the transition back to the pristine state induced by oxygen.7 A Keithley 2400 source-measure unit is used for current-voltage (J-V) characterization. A Solartron SI 1260 is used for impedance spectroscopy with the sample in vacuum. Because of the high conductance under forward bias voltages, capacitance could not be measured reliably under these conditions. Temperature dependent J-V measurements are performed in a vacuum cryostat In all cases, positive bias voltage © 2007 American Chemical Society

Letters

Figure 1. ZnO:PS device with different ligands added are characterized by J-V in the pristine state (a) and after UV illumination (b). Ligands added: none (]), n-propylamine (0), 1-octylaniline (3), 2-naphthalenethiol (]), 1-dodecanethiol (O), and 1-hexadecylamine (4). Inset: schematic layout of the device structure.

refers to the PEDOT:PSS electrode being charged positive with respect to the Al electrode. Results and Discussion Figure 1a shows the J-V characteristics of the ZnO-polymer MIM devices in their pristine, unformed state. Without any added ligands, the device shows diode characteristics when swept between -2 and +2 V, consistent with earlier reports. In forward bias, the current density shows a sharp rise at a bias voltage of 0.5 V and scales as J ∝ V2 for voltages exceeding 0.5 V (Figure 2b). This is consistent with the current being limited by space charge constraints in a layer in the device that is depleted of mobile charge carriers at zero bias. This can be interpreted in terms of the excess electrons of the n type ZnO being trapped by oxygen bound to the surface of the nanoparticle.7 UV illumination triggers the release of oxygen molecules, and this removal of trap sites results in a higher concentration of mobile electrons, allowing for much higher current densities with Ohmic behavior (Figure 1b).7 When adding n-propylamine or 1-octylaniline as an additional coordinating ligand to the ZnO particles, the diode behavior and the persistent photodoping can still be observed, indicating that these ligands do not block the transport of electrical charges when bound to the particles. In contrast, with 2-naphthalenethiol, 1-dodecanethiol, and 1-hexadecylamine as ligands, the current densities are much lower both before and after illumination with UV light. This indicates that these ligands isolate the nanoparticles electrically from one another, blocking any charge transport through the MIM. This shows that the ZnO surface chemistry is indeed crucial in determining the electric properties of the blend. In the J-V scans for devices with 1-hexadecyl amine and 1-octyl aniline, a depolarization current at V ) 0 V can be observed which we ascribe to ionic conductivity due to the presence of protonated ligand. For the other mixtures, the

J. Phys. Chem. C, Vol. 111, No. 28, 2007 10151

Figure 2. (a) Current-voltage characteristics (0 f +8 f -8 f 0 V) of a device with n-propylamine capped ZnO np (ZnO:PS:PA ) 2:4:1 by weight) in the pristine state (O) and of a device with no ligands added to the ZnO np (ZnO:PS ) 1:2 by weight) in the pristine state (3). Scan speed 170 mV/s. (b) Same as (a) but in double logarithmic representation.

depolarization current is below the detection limit, and therefore, the contribution of ionic conductivity to the observed currents is expected to be very small. When sweeping the voltage over a wider range (0 f +8 f -8 f 0 V, Figure 2), we find that the diodes with ZnO:PA show a larger hysteresis in the J-V characteristic than devices without additional ligand. For the latter devices, the J-V trace recorded with increasing and decreasing bias voltage overlap. The hysteresis is larger under reverse bias voltages and is associated with negative differential resistance. Application of reverse bias voltages (V < -5 V) reduces the conductivity at small bias voltages, while application of a forward bias voltage induces a higher conduction. As shown earlier, ZnO-MIM diodes without added ligand show a similar hysteresis but only after a forming process induced by application of a forward bias voltage (∼15 V) or by UV illumination. For bulk metal oxide layers, memory effects are observed only after a forming step.1 In contrast, we show here that, by using zinc oxide nanoparticles with an appropriate surface modification, switching effects can be induced without the need of a forming process. Of all ligands investigated, only n-propylamine is found to induce significant electronic memory effects in the ZnO diodes,15 which have been characterized in more detail using impedance spectroscopy. Figure 3 shows the conductance and capacitance recorded at different probe frequencies (10 Hz-1 kHz) using a voltage that is modulated around a quasi constant bias voltage (alternating current amplitude 100 mV). The bias voltage is swept very slowly from 0 f +5 f -5 f 0 V. The conductance, recorded while sweeping the bias (Figure 3b), shows a large hysteresis, which is the largest for reverse bias voltages. In this respect, this result resembles the static J-V characteristics (Figure 2). Consistent with the J-V characteriza-

10152 J. Phys. Chem. C, Vol. 111, No. 28, 2007

Letters

Figure 3. (a) Conductance of the ZnO:PS:PA device (2:4:1 by weight) as a function of bias voltage (scanned from 0 f +5 f -5 f 0 V). The frequencies displayed are 10 Hz (4), 100 Hz (3), and 1 kHz (]). (b) Conductance and (c) capacitance versus frequency at -0.5 V bias. In b and c, each measurement point is taken after a +5 V (0) or a -5 V (O) pulse with 1 s duration. (d) Capacitance versus bias voltage.

tion, we find that application of a forward bias voltage enhances the conductance probed at small bias voltages. After forward bias stress, the conductance is almost symmetrical around the zero bias voltage, indicating the ZnO layer behaves as a conductor. In contrast, application of reverse bias stress strongly reduces the conductance and restores the asymmetry around zero bias, which illustrates that the devices behaves as a diode in this state. The frequency dependence of the conductance in the high and the low conductivity states (induced by respectively -5 V and +5 V bias) has also been measured, applying a constant bias voltage (-0.5 V, Figure 3b). Before each measurement point in the frequency scan, either a +5 V or a -5 V bias voltage pulse of 1 s is applied to ensure that the device is prepared in the correct state. The high conductivity state is characterized by a conductance that is frequency independent up to 10 kHz. The low conductivity state shows frequency dependent conductivity at high frequencies with Y ∝ ω, in agreement with the generally observed behavior in amorphous materials.16 At frequencies below 100 Hz, the conductance shows a frequency dependence that is less steep. Together with the conductance, also, the capacitance can be monitored. In the low conductivity state, the capacitance probed at -0.5 V bias is independent of frequency and approximately equal to the geometrical capacitance. In the high conductivity state, there is an additional contribution to the capacitance at low frequency. A scan of the bias voltage indicates that this process takes place near bias voltages of ∼ -1 V (Figure 3d). The memory effect can also be characterized by its temperature dependence (Figure 4). Lowering the temperature down to 150 K strongly reduces the hysteresis in the J-V plots. This indicates that the transition from the low to the high conductivity state, that from the high to the low conductivity state, or both transitions involve a thermally activated step. In order to distinguish between these possibilities, we have performed pulse measurements in which the stability of the conduction states is probed in time. The conduction level is probed by applying short (100 ms) -0.5 V bias pulses. In between these probe pulses, the device is kept at zero bias. The conduction level in the low conductivity state, prepared by a 5 s, -5 V bias voltage pulse,

Figure 4. (a) J-V characteristic with (0 f +5 f -5 f 0 V) bias voltage sweep of a ZnO:PS:PA device at 295 (0), 250 (O), and 150 K (4). (b) The ratio of the current density at -0.5 V after a 0 f +5 V scan (on) and after 0 f -5 V scan (off) at varying temperature (150 to 295 K). (c) On/off ratio minus 1 between the current densities probed at -0.5 V after a +5 V pulse, 5 s and after -5 V, 5 s at 295 (0), 275 (O), 255 K (4), 225 K (3), and 190 K (]). (d) Same as c but normalized to unit differential current density at t ) 0.

did not show any particular time dependence in the temperature range studied and neither did the lower conductivity level reached, vary significantly with temperature. In contrast, the conduction level in the high conductivity state, prepared by a 5 s, +5 V pulse, decreases with time. At 295 K, the on/off ratio shows a decay with a characteristic time on the order of 10 s (Figure 4b). Upon lowering the temperature, the spontaneous relaxation for high to the low conductivity becomes much slower, and therefore, this step must be thermally activated. In addition, we find that the conduction level reached after application of the +5 V pulse is temperature dependent, so that also the bias voltage induced transition from the low to the high conductivity state is thermally activated. The on/off ratio is decreases at lower temperature when write and erase pulses of +5 V and -5 V are used. Presumably, the devices can still be switched at lower temperature when using higher voltages. Field dependent activation energies are commonly observed in disordered systems. The role of PEDOT:PSS is not yet clear; however, devices without additional ligand show no reversible switching when the PEDOT:PSS layer is not included in the device structure.9 In conclusion, we have shown that the electronic memory effect in a nanostructured metal oxide can be influenced by surface modification of the particles with coordinating ligands. The mechanism(s) responsible for the memory effects is(are) not yet known, but here, we have shown that chemical modification of ZnO nanoparticles with coordinating ligands opens new opportunities to study and control switching effects in organic-inorganic hybrid materials. Acknowledgment. The work of F.V. forms part of the research programme of the Dutch Polymer Institute (DPI), Project No. DPI#523. We thank Dr. M. Kemerink for his help with the impedance spectroscopy.

Letters Supporting Information Available: Influence of the scan speed and ligand type on the hysteresis, stability of the high and low conductivity state, pulse measurements, dielectric loss and susceptibility graph, and IR transmission and photoluminescence. This material is available free of charge via the Internet at http://pubs.acs.org. References and Notes (1) (a) Thurstans, R. E.; Oxley, D. P. J. Phys. D: Appl. Phys. 2002, 35, 802. (b) Pagnia, H.; Sotnik, N. Phys. Status Solidi A 1988, 108, 11. (c) Dearnaley, G.; Stoneham, A. M.; Morgan, D. V. Rep. Prog. Phys. 1970, 33, 1129. (d) Hickmott, T. W. J. Appl. Phys. 2000, 88, 2805. (2) (a) Beck, A.; Bednorz, J. G.; Gerber, C.; Rossel, C.; Widmer, D. Appl. Phys. Lett. 2000, 77, 139. (b) Baikalov, A.; Wang, Y. Q.; Shen, B.; Lorenz, B.; Tsui, S.; Sun, Y. Y.; Xue, Y. Y.; Chu, C. W. Appl. Phys. Lett. 2003, 83, 957. (c) Szot, K.; Speier, W.; Bihlmayer, G.; Waser, R. Nature Mater 2006, 5, 312. (d) Kinoshita, K.; Tamura, T.; Aoki, M.; Sugiyama, Y.; Tanaka, H. Appl. Phys. Lett. 2006, 89, 103509. (e) Kim, D. C.; Seo, S.; Ahn, S. E.; Suh, D.-S.; Lee, M. J.; Park, B.-H.; Yoo, I. K.; Baek, I. G.; Kim, H.-J.; Yim, E. K.; Lee, J. E.; Park, S. O.; Kim, H. S.; Chung, U.-I.; Moon, J. T.; Ryu, B. I. Appl. Phys. Lett. 2006, 88, 202102. (3) Rozenberg, M. J.; Inoue, I. H.; Sa´nchez, M. J. Phys. ReV. Lett. 2004, 92, 178302. (4) (a) Ralph, J. E.; Woodcock, J. M. J. Non-Cryst. Solids 1972, 7, 236. (b) Kim, D. S.; Kim, Y. H.; Lee, C. E.; Kim, Y. T. Phys. ReV. B 2006, 74, 174430. (c) Simmons, J. G.; Verderber, R. R. Proc. R Soc. London, Ser. A 1967, 301, 77. (d) Brauer, S.; Pagnia, H.; Sotnik, N. Int. J. Electronics 1994, 76, 707. (e) Jeong, D. S.; Choi, B. J.; Hwang, C. S. J. Appl. Phys. 2006, 100, 113724. (f) Tsui, S.; Wang, Y. Q.; Xue, Y. Y.; Chu, C. W. Appl. Phys. Lett. 2006, 89, 123502. (5) Co¨lle, M.; Bu¨chel, M.; de Leeuw, D. M. Org. Electron. 2006, 7, 305.

J. Phys. Chem. C, Vol. 111, No. 28, 2007 10153 (6) (a) Scott, J. C.; Bozano, L. D. AdV. Mater 2007, DOI: 10.1002/ adma.200602564. (b) Mohanta, K.; Majee, S. K.; Batabyal, S. K.; Pal, A. J. J. Phys. Chem. B 2006, 110, 18231. (c) Tseng, R. J.; Baker, C. O.; Shedd, B.; Huang, J. X.; Kaner, R. B.; Ouyang, J. Y.; Yang, Y. Appl. Phys. Lett. 2007, 90, 053101. (7) Verbakel, F.; Meskers, S. C. J.; Janssen, R. A. J. Appl. Phys. Lett. 2006, 89, 102103. (8) Hada, T.; Wasa, K.; Hayakawa, S. Jpn. J. Appl. Phys. 1971, 10, 521. (9) Hunter, O., Jr.; Schaefer, J. A. U.S. Patent 4472296, 1984. (10) (a) Pesika, N. S.; Hu, Z.; Stebe, K. J.; Searson, P. C. J. Phys. Chem. B 2002, 106, 6985. (b) Wong, E. M.; Hoertz, P. G.; Liang, C. J.; Shi, B.-M.; Meyer, G. J.; Searson, P. C. Langmuir 2001, 17, 8362. (11) (a) Sakohara, S.; Ishida, M.; Anderson, M. A. J. Phys. Chem. B 1998, 102, 10169. (b) Cozzoli, P. D.; Kornowski, A.; Weller, H. J. Phys. Chem. B 2005, 109, 2638. (12) (a) Kahn, M. L.; Monge, M.; Colliere, V.; Senocq, F.; Maisonnat, A.; Chaudret, B. AdV. Funct. Mater. 2005, 15, 3. (b) Shim, M.; GuyotSionnest, P. J. Am. Chem. Soc. 2001, 123, 11651. (c) Norberg, N. S.; Gamelin, D. R. J. Phys. Chem. B 2005, 109, 20810. (13) (a) Subramanian, V.; Frechet, J. M. J.; Chang, P. C.; Huang, D. C.; Lee, J. B.; Molesa, S. E.; Murphy, A. R.; Redinger, D. R.; Volkman, S. K. Proc. IEEE 2005, 93, 1330. (b) Sun, B.; Sirringhaus, H. J. Am. Chem. Soc. 2006, 128, 16231. (c) Beek, W. J. E. Thesis, Eindhoven University of Technology, 2005. (d) Ong, B. S.; Li, C.; Li, Y.; Wu, Y.; Loutfy, R. J. Am. Chem. Soc. 2007, 129, 2750. (14) Beek, W. J. E.; Wienk, M. M.; Kemerink, M.; Yang, X.; Janssen, R. A. J. J. Phys. Chem. B 2005, 109, 9505. (15) See Supporting Information. (16) Jonscher, A. K. Dielectric relaxation in solids, Chelsea Dielectrics Press: London, 1983.