Next Article in Journal
Effect of Loading Direction on Tensile-Compressive Mechanical Behaviors of Mg-5Zn-2Gd-0.2Zr Alloy with Heterogeneous Grains
Previous Article in Journal
Effect of Zn on Microstructure and Wear Resistance of Sn-Based Babbitt Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Threshold Switching and Resistive Switching in SnO2-HfO2 Laminated Ultrathin Films

Institute of Physics, University of Tartu, W. Ostwaldi 1, 50411 Tartu, Estonia
*
Author to whom correspondence should be addressed.
Crystals 2024, 14(10), 909; https://doi.org/10.3390/cryst14100909
Submission received: 5 October 2024 / Revised: 15 October 2024 / Accepted: 18 October 2024 / Published: 19 October 2024
(This article belongs to the Section Polycrystalline Ceramics)

Abstract

:
Polycrystalline SnO2-HfO2 nanolaminated thin films were grown by atomic layer deposition (ALD) on SiO2/Si(100) and TiN substrates at 300 °C. The samples, when evaluated electrically, exhibited bipolar resistive switching. The sample object with a stacked oxide layer structure of SnO2 | HfO2 | SnO2 | HfO2 additionally exhibited bidirectional threshold resistive switching properties. The sample with an oxide layer structure of HfO2 | SnO2 | HfO2 displayed bipolar resistive switching with a ratio of high and low resistance states of three orders of magnitude. Endurance tests revealed distinguishable differences between low and high resistance states after 2500 switching cycles.

1. Introduction

Combinations of functional metal oxide films with highly different physical properties, such as conductivity, polarizability, hardness, etc., represented by, e.g., oxides of hafnium and tin, may offer opportunities to expand the usage of functional layers in devices, alongside the possibility of observing interesting properties which may emerge in material stacks. Thereby, HfO2 thin films have a wide range of applications, including in optical coatings [1,2,3], gas sensors [4,5,6], waveguides [2,7], and gate dielectrics [8,9,10]. HfO2 thin films have also been studied from the perspective of memory materials, specifically for use in resistive switching [11,12,13] or ferroelectric memory [14,15,16] devices. SnO2 is a compound material which, in its thin film form, is of considerable interest owing to its well-defined structure and useful physical properties [17]. SnO2 thin films also have many applications, some of which overlap with the applications of HfO2, such as in gas sensors [18,19,20] and waveguides [21,22]. SnO2 films have been studied as transparent conductors [23] and magnetic thin films [24]. In some cases, SnO2 films have also been described as resistive switching materials [25,26,27].
SnO2 and HfO2 thin films have been prepared by sol–gel spin-coating [13,28], pulsed laser deposition [29,30], electron beam evaporation [31,32], reactive sputtering [33,34], spray pyrolysis [5,35], and chemical vapor deposition [36,37]. Atomic layer deposition (ALD) of SnO2 has been studied extensively and realized by many chemical routes, listed in one comprehensive review [38]. The same can be said about HfO2, which has been deposited using a number of different precursor combinations [39].
When it comes to discussing the fabrication methods of specifically resistive switching films, the behavior of SnO2 as a switching media in memristive devices has been studied in few cases, while the deposition method chosen to date has, obviously, been sputtering [40,41,42]. Thereby, another compound, otherwise well known as a high-permittivity dielectric gate oxide, HfO2, is also recognized as probably one of the most widely studied resistive switching materials, and it can be synthesized using a notable variety of techniques before mounting between different electrodes [43]. Specifically, the performance of HfO2-based memristors has been explored via modeling as well [44].
The pairing of SnO2 and HfO2 has also been studied from a few different standpoints. For example, a layer of SnO2 was inserted into HfO2 to act as an oxygen stopping layer [45]. A SnO2 anode was passivated by a HfO2 surface layer [46], and a transistor was made with a SnO2 channel which had a HfO2 gate dielectric on top of it [47].
Methods for fabricating SnO2/HfO2 combinations in electronic devices have not yet included ALD, as far as the authors know. Spin-coating has recently been applied for the deposition of SnO2/HfO2 double layers as channel/gate junctions for field-effect transistors, providing appreciable charge mobility in relatively conductive SnO2 film [48]. Spin-coating accompanying sol–gel synthesis is, however, a technique suited to the growth of rather thick functional films. The aforementioned SnO2/HfO2 junctions for transistor devices had also been fabricated by physical deposition methods [47], though one should account for the suitability of such methods, mainly to two-dimensionally devised structures.
ALD is the thin film growth method most properly suited to the conformal deposition of functional layers on three-dimensional device substrates, e.g., in the case of HfO2 gate dielectrics supporting FinFET transistors [49]. Thereby, thin HfO2 films are considered dielectric insulator material layers applicable together with channel layers made of a two-dimensional crystalline conductor, such as MoS2, on planar [50] as well as on 3D [49] device substrates. In regard to contemporary and prospective concepts of three-dimensional resistive switching memory devices, ALD appears again as the method suited to the synthesis of HfO2-based switching media [51].
The observations made to date imply that combinations of these two physically very different oxide materials may result in attractive properties, and the possible functionality in terms of the appearance of memory effects is worth examination. Prior to the fabrication of 3D device prototypes, the first investigations are to be conducted on solid layers grown onto planar substrates, in order to gain primary knowledge on their structural ordering as well as electrical performance.
The purpose of this work was to use ALD to create a nanolaminated thin film structure consisting of SnO2 and HfO2 layers, and to study it from the standpoint of resistive switching characteristics, after primary structural and compositional evaluation. Since HfO2 and SnO2 both have been reported to exhibit bipolar resistive switching characteristics, as referenced above, a hypothesis was made that layering these different materials together may induce interesting results.

2. Experimental Details

The samples which were studied in this work were deposited in a flow-type ALD reactor [52]. The pressure inside the reactor was 1.8 mbar at room temperature and 2.2 mbar at the working temperature of 300 °C. The precursor for tin was Tin(IV) iodide, SnI4, (99.999%, Sigma Aldrich, St. Louis, MI, USA), which was evaporated at 83 °C inside the reactor, from a glass rod that had an opening at one end. Hafnium precursor was HfCl4, (99.9%, Alfa Aesar, Black Freyer, MA, USA), which was evaporated at 162 °C. Nitrogen, N2 (99.999%, AS Linde Gas, Tallinn, Estonia), was used as the carrier gas and for purging in between precursor pulses. Ozone with a concentration of 220–250 g/m3 was used as an oxidizer, which was produced from O2 (99.999%, AS Linde Gas). All samples were deposited at 300 °C.
The deposition cycle times for SnO2 were 5-2-5-5 s for the following sequence of pulses: metal precursor − N2 purge − O3 − N2 purge. The pulse periods for the HfO2 deposition cycle were 4-3-2-7 s. All films were grown on two different substrates. Firstly, Si(100) substrates were used, having been cleansed and etched prior to the growth. Secondly, for electrical measurements, films were grown on a TiN-covered electrode substrate that had been previously cleansed. The bottom electrode substrates were pieces of Si(100) wafers that had been coated with a crystalline and a conductive 10 nm thick TiN layer, which was deposited via pulsed chemical vapor deposition, using a batch TiCl4/NH3 process [53,54] at temperatures of 450–500 °C in an ASM A412 Large Batch 300 mm reactor at Fraunhofer IPMS-CNT.
The films grown on TiN-covered substrates for electrical measurements were also supplied with Ti/Au electrodes (area 0.204 mm2) that were electron-beam evaporated on top of the films through a shadow mask.
X-ray fluorescence (XRF) spectrometer Rigaku ZSX 400 was used to measure the elemental composition of films, using the in-built software of the ZSX Version 5.55 equipment. A spectroscopic ellipsometer (SE), model GES5-E, was used to measure the thickness of the films. Ellipsometric data were modelled using the Tauc–Lorentz dispersion model. The crystal structure of the samples was evaluated by grazing incidence X-ray diffractometry (GIXRD), using an X-ray diffractometer SmartLab Rigaku. The equipment employs CuKα radiation, which corresponds to an X-ray wavelength of 0.15406 nm.
Cross sections of the laminated structures were studied using scanning transmission electron microscopy (STEM). This was carried out with the FEI Titan Themis 200 instrument with an electron gun operated at 200 kV. The device was equipped with a Bruker SuperX SDD Energy-dispersive X-ray spectroscopy (EDX) system. The EDX system was used for the elemental composition profiling of the SnO2-HfO2 nanolaminates. Electrical measurements were performed by using a Cascade Microtech EPS150TRIAX station, which was controlled by a voltage source and measuring device, namely Keithley 2636A. All electrical measurements were performed at room temperature.

3. Results and Discussion

3.1. Film Growth and Composition

The analysis of the resulting solid films was carried out mainly on samples grown to comparable thicknesses. The number of ALD cycles was kept high enough to ensure a total thickness of the double-layered films of about 20 nm and individual thicknesses of 10 nm for constituent HfO2 and SnO2 layers. This required about 100 deposition cycles for HfO2 and about 50 cycles for SnO2, since the growth rates of reference HfO2 and SnO2 films, measured separately, were 0.11–0.13 and 0.18–0.20 nm/cycle, respectively. The exact growth rate varied somewhat from sample to sample, because the exact growth rate depends on the order of laminates—hafnia grows differently on tin oxide compared to on the silicon substrate and vice versa—and on the number of deposition cycles. The growth rate of HfO2, 0.1 nm/cycle, corresponded to that which was previously reported in the literature at the same temperature and using the same precursor chemistry [55]. The growth rate of SnO2 was harder to compare with the values reported in the literature, since the ALD process based on SnI4 and O3 precursors has previously been reported only by our group [56]. In the present study, all resulting films possessed thicknesses in the range of 22–25 nm.
Since the precursors used contained I and Cl, the amounts of these elements in the films were assessed as well. Since the optimization of growth for both HfO2 and SnO2 had already been published elsewhere [55,56], the amounts of these residues could also be held at low values in the present study, as the content of I was <1.0 at% and the content of Cl was <0.1 at%.

3.2. Film Structure

When the constituent oxide layer thickness was less than 5 nm, no distinct diffraction maxima were observed on the diffractogram (top pattern in Figure 1) of a four-layer laminate. Constituent layer thicknesses exceeding 10 nm enabled the appearance of a well-defined set of maxima, which allowed one to recognize the formation and presence of the monoclinic HfO2 and a mixture of cubic/tetragonal SnO2 (bottom two patterns in Figure 1).
Comparisons of STEM images and the EDX composition profiling of the 25 × SnO2 + 50 × HfO2 + 25 × SnO2 + 50 × HfO2 sample demonstrate that a multilayer, i.e., nanolaminate structure has been formed (Figure 2). The appearance of distinguishable lattice planes is indicative of a nanocrystalline order, which is not clearly observed in the GIXRD diffractogram.

3.3. Electrical Properties

Resistive switching as a phenomenon manifests itself via the alternate stabilization of high and low resistivity states in the solid material layer. The persistence of such memory states is not necessarily flawless. One of the drawbacks is current leakage through low resistive state (LRS) cells in the memory crossbar device, i.e., the sneak path problem [57]. Usually, a selector device like a transistor can be used to restrict sneak path currents; however, simpler two-terminal selector devices are preferable to maintain the overall minimal footprint of the memory cell [58]. A selector device is a highly non-linear device like a diode or a threshold switch. In the case of a threshold switch, the device provides a high resistance state (HRS) until the applied voltage is increased to a certain value, and then it switches to an LRS. The LRS is retained while the voltage is reduced down to a threshold voltage value. In this way, an LRS forms a window of voltage that can be used to switch a resistive switching device that is connected in series. In an attempt to reduce the risk of contamination by common diffusive contaminants like Cu and Ag in silicon electronics, Sonde et al. [59] studied HfO2-based conductive-bridge random-access memory (CBRAM), replacing the commonly used anode material, e.g., Cu or Ag, with Sb. CBRAM is based on the electrochemical metallization mechanism of resistive switching. The aforementioned study demonstrated the bipolar resistive switching and threshold switching capabilities of Pt/HfO2/Sn structures. Figure 3 shows the current–voltage characteristic of a bidirectional threshold switching cell composed of a resistive switching medium sandwiched between Ti/Au and TiN electrodes. The resistive switching medium was deposited using a cycle sequence of 25 × SnO2 + 50 × HfO2 + 25 × SnO2 + 50 × HfO2, resulting in a multilayered structure consisting of 5 nm thick SnO2, 7 nm thick HfO2, 5 nm thick SnO2, and 7 nm thick HfO2 layers. Similar electrical behavior was observed by Sonde et al. [59] in a 4 nm thick HfO2 device with Pt and Sn electrodes. Interestingly, in our study, the switching voltages at negative polarity coincide with the aforementioned study, whereas the asymmetrical I-V characteristics in our case may be related to the asymmetrical structure of the film [59]. In an LRS, the operating voltage window for a series device ranged approximately from −1 V down to −3 V in the case of negative voltage polarity, and from 1 V to nearly 2 V at positive voltage polarity. The ratio between current values in the HRS and LRS increased by over three orders of magnitude, whereas Sonde et al. [59] achieved approximately one order of magnitude.
Figure 4 shows the I-V curve of a bipolar resistive switching device containing a solid medium deposited using a cycle sequence of 70 × HfO2 + 30 × SnO2 + 70 × HfO2. The ratio between the HRS and LRS was, again, measured to be over around three orders of magnitude (Figure 4a). Nevertheless, one can notice that the I-V curve is rather unstable; thus, optimizing switching voltage pulses for endurance measurement would need further improvement. This device withstood about 2500 switching cycles, noticeably demonstrating LRS currents below 10 nA. Thereafter, the distinguishable difference between the HRS and LRS collapsed due to the resistivity loss in the HRS (Figure 4b). It is plausible that at such low operating currents, the source-meter current compliance was not sufficiently fast, causing the degradation of the HRS and the noticeably smaller HRS to LRS ratio. As seen in the discussion above, two specific samples of laminated structures exhibited resistive switching characteristics. Other systems were examined as well, varying the number of layers and their order, but they did not exhibit publication-worthy resistive switching characteristics. It is plausible that a certain balance between the more conductive SnO2 and the more insulating HfO2 is necessary for successful switching properties, and the reported films had this balance, but an exact assessment of what this balance entails is beyond the scope of this work and should be left for further studies.

4. Summary

Nanolaminates consisting of SnO2 and HfO2 layers were deposited by ALD, using SnI4, HfCl4, and O3 as precursors. SnO2 was found to grow in its cubic or tetragonal phase, whereas HfO2 was formed mostly in its monoclinic phase. Films containing relatively thin layers, grown to thicknesses below 7–8 nm, remained weakly crystallized. A film deposited using a cycle sequence of 25 × SnO2 + 50 × HfO2 + 25 × SnO2 + 50 × HfO2 and mounted between Au and TiN electrodes exhibited threshold switching characteristics, with the ratio between the HRS and LRS increasing by nearly three orders of magnitude. The film deposited using a cycle sequence of 70 × HfO2 + 30 × SnO2 + 70 × HfO2 acted as a bipolar resistive switching device, with a ratio between the HRS and LRS of around three orders of magnitude. The latter device withstood about 2500 cycles before failure.

Author Contributions

Conceptualization, K.K. (Kristjan Kalam) and K.K. (Kaupo Kukli); Data curation, K.K. (Kristjan Kalam); Formal analysis, M.-E.A., J.M. and M.O.; Funding acquisition, K.K. (Kristjan Kalam) and K.K. (Kaupo Kukli); Investigation, M.-E.A., J.M., M.O. and P.R.; Methodology, K.K. (Kristjan Kalam), M.-E.A., J.M. and M.O.; Project administration, K.K. (Kristjan Kalam) and K.K. (Kaupo Kukli); Resources, K.K. (Kristjan Kalam) and K.K. (Kaupo Kukli); Supervision, K.K. (Kristjan Kalam) and K.K. (Kaupo Kukli); Validation, K.K. (Kristjan Kalam); Writing—original draft, K.K. (Kristjan Kalam); Writing—review and editing, J.M., M.O. and K.K. (Kaupo Kukli). All authors have read and agreed to the published version of the manuscript.

Funding

This study was partially supported by the European Regional Development Fund project “Emerging orders in quantum and nanomaterials” (TK134), and the Estonian Research Agency (PRG753, PUTJD1220).

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Khoshman, J.M.; Khan, A.; Kordesch, M.E. Amorphous hafnium oxide thin films for antireflection optical coatings. Surf. Coat. Technol. 2008, 202, 2500–2502. [Google Scholar] [CrossRef]
  2. Khoshman, J.M.; Kordesch, M.E. Optical properties of a-HfO2 thin films. Surf. Coat. Technol. 2006, 201, 3530–3535. [Google Scholar] [CrossRef]
  3. Wiatrowski, A.; Obstarczyk, A.; Mazur, M.; Kaczmarek, D.; Wojcieszak, D. Characterization of HfO2 optical coatings deposited by MF magnetron sputtering. Coatings 2019, 9, 106. [Google Scholar] [CrossRef]
  4. Capone, S.; Leo, G.; Rella, R.; Siciliano, P.; Vasanelli, L.; Alvisi, M.; Mirenghi, L.; Rizzo, A. Physical characterization of hafnium oxide thin films and their application as gas sensing devices. J. Vac. Sci. Technol. A 1998, 16, 3564–3568. [Google Scholar] [CrossRef]
  5. Avila-García, A.; García-Hipólito, M. Characterization of gas sensing HfO2 coatings synthesized by spray pyrolysis technique. Sens. Actuators B Chem. 2008, 133, 302–307. [Google Scholar] [CrossRef]
  6. Karaduman, I.; Acar, S. The gas sensing properties of hafnium oxide thin films depending on the annealing environment. Mod. Phys. Lett. B 2017, 31, 1750284. [Google Scholar] [CrossRef]
  7. Tiwari, K.; Sharma, S.C.; Hozhabri, N. Hafnium dioxide as a dielectric for highly-sensitive waveguide-coupled surface plasmon resonance sensors. Aip Adv. 2016, 6, 045217. [Google Scholar] [CrossRef]
  8. He, G.; Zhu, L.Q.; Liu, M.; Fang, Q.; Zhang, L.D. Optical and electrical properties of plasma-oxidation derived HfO2 gate dielectric films. Appl. Surf. Sci. 2007, 253, 3413–3418. [Google Scholar] [CrossRef]
  9. Gusev, E.P.; Cabral Jr, C.; Copel, M.; D’emic, C.; Gribelyuk, M. Ultrathin HfO2 films grown on silicon by atomic layer deposition for advanced gate dielectrics applications. Microelectron. Eng. 2003, 69, 145–151. [Google Scholar] [CrossRef]
  10. Huang, A.P.; Yang, Z.C.; Chu, P.K. Advances in Solid State Circuits Technologies; InTech: Rijeka, Croatia, 2010; Hafnium-based high-k gate dielectrics; pp. 333–350. ISBN 978-953-307-086-5. [Google Scholar]
  11. Wang, Y.; Liu, Q.; Long, S.; Wang, W.; Wang, Q.; Zhang, M.; Zhang, S.; Li, Y.; Zuo, Q.; Yang, J.; et al. Investigation of resistive switching in Cu-doped HfO2 thin film for multilevel non-volatile memory applications. Nanotechnology 2009, 21, 045202. [Google Scholar] [CrossRef]
  12. Liu, C.F.; Tang, X.G.; Wang, L.Q.; Tang, H.; Jiang, Y.P.; Liu, Q.X.; Li, W.H.; Tang, Z.H. Resistive switching characteristics of HfO2 thin films on mica substrates prepared by sol-gel process. Nanomaterials 2019, 9, 1124. [Google Scholar] [CrossRef] [PubMed]
  13. Ramadoss, A.; Krishnamoorthy, K.; Kim, S.J. Resistive switching behaviors of HfO2 thin films by sol–gel spin coating for nonvolatile memory applications. Appl. Phys. Express 2012, 5, 085803. [Google Scholar] [CrossRef]
  14. Mueller, S.; Mueller, J.; Singh, A.; Riedel, S.; Sundqvist, J.; Schroeder, U.; Mikolajick, T. Incipient ferroelectricity in Al-doped HfO2 thin films. Adv. Funct. Mater. 2012, 22, 2412–2417. [Google Scholar] [CrossRef]
  15. Sang, X.; Grimley, E.D.; Schenk, T.; Schroeder, U.; LeBeau, J.M. On the structural origins of ferroelectricity in HfO2 thin films. Appl. Phys. Lett. 2015, 106, 162905. [Google Scholar] [CrossRef]
  16. Böscke, T.S.; Müller, J.; Bräuhaus, D.; Schröder, U.; Böttger, U.J.A.P.L. Ferroelectricity in hafnium oxide thin films. Appl. Phys. Lett. 2011, 99, 102903. [Google Scholar] [CrossRef]
  17. Singha, K.K.; Singh, P.P.; Narzary, R.; Mondal, A.; Gupta, M.; Sathe, V.G.; Kumar, D.; Srivastava, S.K. Crystal Structure, Raman Spectroscopy and Optical Property Study of Mg-Doped SnO2 Compounds for Optoelectronic Devices. Crystals 2023, 13, 932. [Google Scholar] [CrossRef]
  18. Sberveglieri, G. Classical and novel techniques for the preparation of SnO2 thin-film gas sensors. Sens. Actuators B Chem. 1992, 6, 239–247. [Google Scholar] [CrossRef]
  19. Oyabu, T. Sensing characteristics of SnO2 thin film gas sensor. J. Appl. Phys. 1982, 53, 2785–2787. [Google Scholar] [CrossRef]
  20. Lee, S.W.; Tsai, P.P.; Chen, H. Comparison study of SnO2 thin-and thick-film gas sensors. Sens. Actuators B Chem. 2000, 67, 122–127. [Google Scholar] [CrossRef]
  21. Berneschi, S.; Bhaktha, S.N.B.; Chiappini, A.; Chiasera, A.; Ferrari, M.; Kinowski, C.; Turrell, S.; Trono, C.; Brenci, M.; Cacciari, I.; et al. Highly photorefractive Eu3+ activated sol-gel SnO2-SnO2 thin film waveguides. Integr. Opt. Devices Mater. Technol. 2010, 7604, 246–251. [Google Scholar] [CrossRef]
  22. Tran, T.N.L.; Armellini, C.; Varas, S.; Carpentiero, A.; Chiappini, A.; Głuchowski, P.; Iacob, E.; Ferrari, M. Assessment of SnO2-Nanocrystal-Based Luminescent Glass-Ceramic Waveguides for Integrated Photonics. Ceram. Int. 2021, 47, 5534–5541. [Google Scholar] [CrossRef]
  23. Qadri, S.B.; Kim, H.; Khan, H.R.; Piqué, A.; Horwitz, J.S.; Chrisey, D.; Kim, W.J.; Skelton, E.F. Transparent Conducting Films of In2O3–ZrO2, SnO2–ZrO2 and ZnO–ZrO2. Thin Solid Films 2000, 377, 750–754. [Google Scholar] [CrossRef]
  24. Manjula, N.; Selvan, G. Magnetic and Antibacterial Properties of Zr-Doped SnO2 Nanopowders. J. Mater. Sci. Mater. Electron. 2017, 28, 15056–15064. [Google Scholar] [CrossRef]
  25. Nagashima, K.; Yanagida, T.; Oka, K.; Kawai, T. Unipolar Resistive Switching Characteristics of Room Temperature Grown SnO2 Thin Films. Appl. Phys. Lett. 2009, 94, 242902. [Google Scholar] [CrossRef]
  26. Almeida, S.; Aguirre, B.; Marquez, N.; McClure, J.; Zubia, D. Resistive Switching of SnO2 Thin Films on Glass Substrates. Integr. Ferroelectr. 2011, 126, 117–124. [Google Scholar] [CrossRef]
  27. Kalateh, A.; Jalali, A.; Kamali Ashtiani, M.J.; Mohammadimasoudi, M.; Bastami, H.; Mohseni, M. Resistive Switching Transparent SnO2 Thin Film Sensitive to Light and Humidity. Sci. Rep. 2023, 13, 20036. [Google Scholar] [CrossRef]
  28. Ganchev, M.; Katerski, A.; Stankova, S.; Eensalu, J.S.; Terziyska, P.; Gergova, R.; Popkirov, G.; Vitanov, P. Spin-Coating of SnO2 Thin Films. J. Phys. Conf. Ser. 2019, 1186, 012027. [Google Scholar] [CrossRef]
  29. Kim, H.; Auyeung, R.C.Y.; Piqué, A. Transparent Conducting F-Doped SnO2 Thin Films Grown by Pulsed Laser Deposition. Thin Solid Films 2008, 516, 5052–5056. [Google Scholar] [CrossRef]
  30. Luo, Y.; Tang, Z.; Yin, X.; Chen, C.; Fan, Z.; Qin, M.; Zeng, M.; Zhou, Z.; Hu, X.; Gao, X. Ferroelectricity in Dopant-Free HfO2 Thin Films Prepared by Pulsed Laser Deposition. J. Mater. 2022, 8, 311–318. [Google Scholar] [CrossRef]
  31. Alnuwaiser, M.A.; Javaid, K.; Jacob, J.; Saleem, M.; Liang, L.; Cao, H.; Amami, M.; Li, Y.; Li, D. Annealing Induced Morphology Evolution and Phase Transition in SnOx Thin Films Grown by E-Beam Evaporation Method. Inorg. Chem. Commun. 2022, 140, 109473. [Google Scholar] [CrossRef]
  32. Alvisi, M.; De Tomasi, F.; Perrone, M.R.; Protopapa, M.L.; Rizzo, A.; Sarto, F.; Scaglione, S. Laser Damage Dependence on Structural and Optical Properties of Ion-Assisted HfO2 Thin Films. Thin Solid Films 2001, 396, 44–52. [Google Scholar] [CrossRef]
  33. Gubbins, M.A.; Casey, V.; Newcomb, S.B. Nanostructural Characterisation of SnO2 Thin Films Prepared by Reactive RF Magnetron Sputtering of Tin. Thin Solid Films 2002, 405, 270–275. [Google Scholar] [CrossRef]
  34. Martínez, F.L.; Toledano-Luque, M.; Gandía, J.J.; Cárabe, J.; Bohne, W.; Röhrich, J.; Strub, E.; Mártil, I. Optical Properties and Structure of HfO2 Thin Films Grown by High Pressure Reactive Sputtering. J. Phys. D Appl. Phys. 2007, 40, 5256. [Google Scholar] [CrossRef]
  35. Gordillo, G.; Moreno, L.C.; De la Cruz, W.; Teheran, P. Preparation and Characterization of SnO2 Thin Films Deposited by Spray Pyrolysis from SnCl2 and SnCl4 Precursors. Thin Solid Films 1994, 252, 61–66. [Google Scholar] [CrossRef]
  36. Jeong, J.; Choi, S.-P.; Chang, C.I.; Shin, D.C.; Park, J.S.; Lee, B.T.; Park, Y.-J.; Song, H.-J. Photoluminescence Properties of SnO2 Thin Films Grown by Thermal CVD. Solid State Commun. 2003, 127, 595–597. [Google Scholar] [CrossRef]
  37. Park, J.; Park, B.K.; Cho, M.; Hwang, C.S.; Oh, K.; Yang, D.Y. Chemical Vapor Deposition of HfO2 Thin Films Using a Novel Carbon-Free Precursor: Characterization of the Interface with the Silicon Substrate. J. Electrochem. Soc. 2001, 149, G89. [Google Scholar] [CrossRef]
  38. Nazarov, D.V.; Bobrysheva, N.P.; Osmolovskaya, O.M.; Osmolovsky, M.G.; Smirnov, V.M. Atomic Layer Deposition of Tin Dioxide Nanofilms: A Review. Rev. Adv. Mater. Sci. 2015, 40, 262–275. [Google Scholar]
  39. Choi, J.H.; Mao, Y.; Chang, J.P. Development of Hafnium-Based High-k Materials—A Review. Mater. Sci. Eng. R Rep. 2011, 72, 97–136. [Google Scholar] [CrossRef]
  40. Ismail, M.; Mahata, C.; Kang, M.; Kim, S. SnO2-Based Memory Device with Filamentary Switching Mechanism for Advanced Data Storage and Computing. Nanomaterials 2023, 13, 2603. [Google Scholar] [CrossRef]
  41. Yun, M.J.; Kim, K.H.; Bea, D.; Jung, J.; Kim, S.; Kim, H.-D. Improved Resistive Switching of SnO2-Based Resistive Random Access Memory Devices Using Post Microwave Treatment. J. Electr. Eng. Technol. 2021, 16, 1011–1017. [Google Scholar] [CrossRef]
  42. Dastgeer, G.; Afzal, A.M.; Aziz, J.; Hussain, S.; Jaffery, S.H.A.; Kim, D.; Imran, M.; Assiri, M.A. Flexible Memory Device Composed of Metal-Oxide and Two-Dimensional Material (SnO2/WTe2) Exhibiting Stable Resistive Switching. Materials 2021, 14, 7535. [Google Scholar] [CrossRef] [PubMed]
  43. Banerjee, W.; Kashir, A.; Kamba, S. Hafnium Oxide (HfO2)—A Multifunctional Oxide: A Review on the Prospect and Challenges of Hafnium Oxide in Resistive Switching and Ferroelectric Memories. Small 2022, 18, 2107575. [Google Scholar] [CrossRef] [PubMed]
  44. Lin, J.; Liu, H.; Wang, S.; Zhang, S. Modeling and Simulation of Hafnium Oxide RRAM Based on Oxygen Vacancy Conduction. Crystals 2021, 11, 1462. [Google Scholar] [CrossRef]
  45. Chen, W.-C.; Qin, S.; Yu, Z.; Wong, H.-S.P. Reduced HfO2 Resistive Memory Variability by Inserting a Thin SnO2 as Oxygen Stopping Layer. IEEE Electron Device Lett. 2021, 42, 1778–1781. [Google Scholar] [CrossRef]
  46. Yesibolati, N.; Shahid, M.; Chen, W.; Hedhili, M.N.; Reuter, M.C.; Ross, F.M.; Alshareef, H.N. SnO2 Anode Surface Passivation by Atomic Layer Deposited HfO2 Improves Li-Ion Battery Performance. Small 2014, 10, 2849–2858. [Google Scholar] [CrossRef]
  47. Yen, T.J.; Chin, A.; Gritsenko, V. High-Performance Top-Gate Thin-Film Transistor with an Ultra-Thin Channel Layer. Nanomaterials 2020, 10, 2145. [Google Scholar] [CrossRef]
  48. Avis, C.; Billah, M.M.; Kim, Y.G.; Siddik, A.B.; Jang, J. Analysis of the Solution-Processed a-SnOx and HfO2 Interface for Applications in Thin-Film Transistors. ACS Appl. Electron. Mater. 2021, 3, 651–657. [Google Scholar] [CrossRef]
  49. Chen, M.-L.; Sun, X.; Liu, H.; Wang, H.; Zhu, Q.; Wang, S.; Du, H.; Dong, B.; Zhang, J.; Sun, Y.; et al. A FinFET with One Atomic Layer Channel. Nat. Commun. 2020, 11, 1205. [Google Scholar] [CrossRef]
  50. Schilirò, E.; Panasci, S.E.; Mio, A.M.; Nicotra, G.; Agnello, S.; Pecz, B.; Radnoczi, G.Z.; Deretzis, I.; La Magna, A.; Roccaforte, F.; et al. Direct Atomic Layer Deposition of Ultra-Thin Al2O3 and HfO2 Films on Gold-Supported Monolayer MoS2. Appl. Surf. Sci. 2023, 630, 157476. [Google Scholar] [CrossRef]
  51. Li, Y.; Tang, J.; Gao, B.; Yao, J.; Fan, A.; Yan, B.; Yang, Y.; Xi, Y.; Li, Y.; Li, J.; et al. Monolithic Three-Dimensional Integration of RRAM-Based Hybrid Memory Architecture for One-Shot Learning. Nat. Commun. 2023, 14, 7140. [Google Scholar] [CrossRef]
  52. Arroval, T.; Aarik, L.; Rammula, R.; Kruusla, V.; Aarik, J. Effect of Substrate-Enhanced and Inhibited Growth on Atomic Layer Deposition and Properties of Aluminum–Titanium Oxide Films. Thin Solid Films 2016, 600, 119–125. [Google Scholar] [CrossRef]
  53. Granneman, E.; Fischer, P.; Pierreux, D.; Terhorst, H.; Zagwijn, P. Batch ALD: Characteristics, Comparison with Single Wafer ALD, and Examples. Surf. Coat. Technol. 2007, 201, 8899–8904. [Google Scholar] [CrossRef]
  54. Zagwijn, P.M.; Verweij, W.; Pierreux, D.; Adjeroud, N.; Bankras, R.; Oosterlaken, E.; Snijders, G.J.; van den Hout, M.; Fischer, P.; Wilhelm, R.; et al. Novel Batch Titanium Nitride CVD Process for Advanced Metal Electrodes. ECS Trans. 2008, 13, 459–468. [Google Scholar] [CrossRef]
  55. Aarik, L.; Arroval, T.; Mändar, H.; Rammula, R.; Aarik, J. Influence of Oxygen Precursors on Atomic Layer Deposition of HfO2 and Hafnium-Titanium Oxide Films: Comparison of O3− and H2O-Based Processes. Appl. Surf. Sci. 2020, 530, 147229. [Google Scholar] [CrossRef]
  56. Kalam, K.; Ritslaid, P.; Käämbre, T.; Tamm, A.; Kukli, K. Properties of Tin Oxide Films Grown by Atomic Layer Deposition from Tin Tetraiodide and Ozone. Beilstein J. Nanotechnol. 2023, 14, 1085–1092. [Google Scholar] [CrossRef]
  57. Lanza, M.; Wong, H.-S.P.; Pop, E.; Ielmini, D.; Strukov, D.; Regan, B.C.; Larcher, L.; Villena, M.A.; Yang, J.J.; Xia, Q.; et al. Recommended Methods to Study Resistive Switching Devices. Adv. Electron. Mater. 2019, 5, 1800143. [Google Scholar] [CrossRef]
  58. Shi, L.; Zheng, G.; Tian, B.; Dkhil, B.; Duan, C. Research Progress on Solutions to the Sneak Path Issue in Memristor Crossbar Arrays. Nanoscale Adv. 2020, 2, 1811–1827. [Google Scholar] [CrossRef]
  59. Sonde, S.; Chakrabarti, B.; Liu, Y.; Sasikumar, K.; Lin, J.; Stan, L.; Divan, R.; Chaudhuri, J.; Sbrockey, N.; Weiss, D.; et al. Silicon Compatible Sn-Based Resistive Switching Memory. Nanoscale 2018, 10, 9441–9449. [Google Scholar] [CrossRef]
Figure 1. GIXRD patterns of HfO2-SnO2 films. Miller indices attributed to the diffraction maxima are indicated on the graph, where T and C correspond to the tetragonal and cubic crystalline phases of SnO2, and M corresponds to the monoclinic phase of HfO2. The layered structure of the films is indicated by labels, wherein the numbers denote the numbers of cycles applied for the deposition of constituent oxide layers.
Figure 1. GIXRD patterns of HfO2-SnO2 films. Miller indices attributed to the diffraction maxima are indicated on the graph, where T and C correspond to the tetragonal and cubic crystalline phases of SnO2, and M corresponds to the monoclinic phase of HfO2. The layered structure of the films is indicated by labels, wherein the numbers denote the numbers of cycles applied for the deposition of constituent oxide layers.
Crystals 14 00909 g001
Figure 2. Cross-sectional STEM HAADF image and EDX composition profiling taken from the lamella that was made of the nanolaminate 25 × SnO2 + 50 × HfO2 + 25 × SnO2 + 50 × HfO2.
Figure 2. Cross-sectional STEM HAADF image and EDX composition profiling taken from the lamella that was made of the nanolaminate 25 × SnO2 + 50 × HfO2 + 25 × SnO2 + 50 × HfO2.
Crystals 14 00909 g002
Figure 3. Current–voltage characteristic of a threshold resistive switching device, measured from a film deposited using a cycle sequence of 25 × SnO2 + 50 × HfO2 + 25 × SnO2 + 50 × HfO2. Upward arrows indicate the SET process and downward arrows indicate the RESET process.
Figure 3. Current–voltage characteristic of a threshold resistive switching device, measured from a film deposited using a cycle sequence of 25 × SnO2 + 50 × HfO2 + 25 × SnO2 + 50 × HfO2. Upward arrows indicate the SET process and downward arrows indicate the RESET process.
Crystals 14 00909 g003
Figure 4. Current–voltage characteristic of a bipolar resistive switching device (a) and endurance graph (b), measured from a film deposited using a cycle sequence of 70 × HfO2 + 30 × SnO2 + 70 × HfO2. HRS and LRS data are marked with black and red circles, respectively. Upward arrows indicate the SET process and downward arrows indicate the RESET process.
Figure 4. Current–voltage characteristic of a bipolar resistive switching device (a) and endurance graph (b), measured from a film deposited using a cycle sequence of 70 × HfO2 + 30 × SnO2 + 70 × HfO2. HRS and LRS data are marked with black and red circles, respectively. Upward arrows indicate the SET process and downward arrows indicate the RESET process.
Crystals 14 00909 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kalam, K.; Aan, M.-E.; Merisalu, J.; Otsus, M.; Ritslaid, P.; Kukli, K. Threshold Switching and Resistive Switching in SnO2-HfO2 Laminated Ultrathin Films. Crystals 2024, 14, 909. https://doi.org/10.3390/cryst14100909

AMA Style

Kalam K, Aan M-E, Merisalu J, Otsus M, Ritslaid P, Kukli K. Threshold Switching and Resistive Switching in SnO2-HfO2 Laminated Ultrathin Films. Crystals. 2024; 14(10):909. https://doi.org/10.3390/cryst14100909

Chicago/Turabian Style

Kalam, Kristjan, Mark-Erik Aan, Joonas Merisalu, Markus Otsus, Peeter Ritslaid, and Kaupo Kukli. 2024. "Threshold Switching and Resistive Switching in SnO2-HfO2 Laminated Ultrathin Films" Crystals 14, no. 10: 909. https://doi.org/10.3390/cryst14100909

APA Style

Kalam, K., Aan, M. -E., Merisalu, J., Otsus, M., Ritslaid, P., & Kukli, K. (2024). Threshold Switching and Resistive Switching in SnO2-HfO2 Laminated Ultrathin Films. Crystals, 14(10), 909. https://doi.org/10.3390/cryst14100909

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop