Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Ferroelectric transistors based on shear-transformation-mediated rhombohedral-stacked molybdenum disulfide

Abstract

To develop low-power, non-volatile computing-in-memory device using ferroelectric transistor technologies, ferroelectric channel materials with scaled thicknesses are required. Two-dimensional semiconductors, such as molybdenum disulfide (MoS2), equipped with sliding ferroelectricity could provide an answer. However, achieving switchable electric polarization in epitaxial MoS2 remains challenging due to the absence of mobile domain boundaries. Here we show that polarity-switchable epitaxial rhombohedral-stacked (3R) MoS2 can be used as a ferroelectric channel in ferroelectric memory transistors. We show that a shear transformation can spontaneously occur in 3R MoS2 epilayers, producing heterostructures with stable ferroelectric domains embedded in a highly dislocated and unstable non-ferroelectric matrix. This diffusionless phase transformation process produces mobile screw dislocations that enable collective polarity control of 3R MoS2 via an electric field. Polarization–electric-field measurements reveal a switching field of 0.036 V nm−1 for shear-transformed 3R MoS2. Our sliding ferroelectric transistors are non-volatile memory units with thicknesses of only two atomic layers and exhibit an average memory window of 7 V with an applied voltage of 10 V, retention times greater than 104 seconds and endurance greater than 104 cycles.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Epitaxial 3R bilayer MoS2 with switchable electric polarization.
Fig. 2: PFM measurements, the vdW 3R MoS2 capacitor and the ferroelectric property of the ST-3R bilayer MoS2.
Fig. 3: Device characteristics of 3R MoS2 sliding-ferroelectric FETs.

Similar content being viewed by others

Data availability

Source data are provided with this paper. All other data that support the findings in this study are available from the corresponding authors upon reasonable request.

References

  1. Khan, A. I., Keshavarzi, A. & Datta, S. The future of ferroelectric field-effect transistor technology. Nat. Electron. 3, 588–597 (2020).

    Article  Google Scholar 

  2. Mikolajick, T., Slesazeck, S., Park, M. H. & Schroeder, U. Ferroelectric hafnium oxide for ferroelectric random-access memories and ferroelectric field-effect transistors. MRS Bull. 43, 340–346 (2018).

    Article  Google Scholar 

  3. Si, M. et al. A ferroelectric semiconductor field-effect transistor. Nat. Electron. 2, 580–586 (2019).

    Article  Google Scholar 

  4. Miller, S., Nasby, R., Schwank, J., Rodgers, M. & Dressendorfer, P. Device modeling of ferroelectric capacitors. J. Appl. Phys. 68, 6463–6471 (1990).

    Article  Google Scholar 

  5. Ni, K., Jerry, M., Smith, J. A. & Datta, S. A circuit compatible accurate compact model for ferroelectric-FETs. In IEEE Symposium on VLSI Technology. 131–132 (IEEE, 2018); https://doi.org/10.1109/VLSIT.2018.8510622

  6. Jiang, Y. et al. Enabling ultra-low-voltage switching in BaTiO3. Nat. Mater. 21, 779–785 (2022).

    Article  Google Scholar 

  7. Cheema, S. S. et al. Enhanced ferroelectricity in ultrathin films grown directly on silicon. Nature 580, 478–482 (2020).

    Article  Google Scholar 

  8. Ahn, C. H., Rabe, K. M. & Triscone, J. M. Ferroelectricity at the nanoscale: local polarization in oxide thin films and heterostructures. Science 303, 488–491 (2004).

    Article  Google Scholar 

  9. Mehta, R. R., Silverman, B. D. & Jacobs, J. T. Depolarization fields in thin ferroelectric films. J. Appl. Phys. 44, 3379–3385 (1973).

    Article  Google Scholar 

  10. Cheema, S. S. et al. Emergent ferroelectricity in subnanometer binary oxide films on silicon. Science 376, 648–652 (2022).

    Article  Google Scholar 

  11. Tan, A. J. et al. Ferroelectric HfO2 memory transistors with high-κ interfacial layer and write endurance exceeding 1010 cycles. IEEE Electron Dev. Lett. 42, 994–997 (2021).

    Article  Google Scholar 

  12. Hoffmann, M. et al. Fast read-after-write and depolarization fields in high endurance n-type ferroelectric FETs. IEEE Electron Dev. Lett. 43, 717–720 (2022).

    Article  Google Scholar 

  13. Yasuda, K., Wang, X., Watanabe, K., Taniguchi, T. & Jarillo-Herrero, P. Stacking-engineered ferroelectricity in bilayer boron nitride. Science 372, 1458–1462 (2021).

    Article  Google Scholar 

  14. Vizner Stern, M. et al. Interfacial ferroelectricity by van der Waals sliding. Science 372, 1462–1466 (2021).

    Article  Google Scholar 

  15. Rogée, L. et al. Ferroelectricity in untwisted heterobilayers of transition metal dichalcogenides. Science 376, 973–978 (2022).

    Article  Google Scholar 

  16. Weston, A. et al. Interfacial ferroelectricity in marginally twisted 2D semiconductors. Nat. Nanotechnol. 17, 390–395 (2022).

    Article  Google Scholar 

  17. Wang, X. et al. Interfacial ferroelectricity in rhombohedral-stacked bilayer transition metal dichalcogenides. Nat. Nanotechnol. 17, 367–371 (2022).

    Article  Google Scholar 

  18. Miao, L. P. et al. Direct observation of geometric and sliding ferroelectricity in an amphidynamic crystal. Nat. Mater. 21, 1158–1164 (2022).

    Article  Google Scholar 

  19. Wang, Q. et al. Layer-by-layer epitaxy of multi-layer MoS2 wafers. Natl Sci. Rev. 9, nwac077 (2022).

    Article  MathSciNet  Google Scholar 

  20. Liu, L. et al. Uniform nucleation and epitaxy of bilayer molybdenum disulfide on sapphire. Nature 605, 69–75 (2022).

    Article  Google Scholar 

  21. Kim, J. H. et al. Interface-driven partial dislocation formation in 2D heterostructures. Adv. Mater. 31, e1807486 (2019).

    Article  Google Scholar 

  22. Zhang, Z. & Lagally, M. G. Atomistic processes in the early stages of thin-film growth. Science 276, 377–383 (1997).

    Article  Google Scholar 

  23. Witten, T. A. & Sander, L. M. Diffusion-limited aggregation. Phys. Rev. B 27, 5686–5697 (1983).

    Article  MathSciNet  Google Scholar 

  24. Weston, A. et al. Atomic reconstruction in twisted bilayers of transition metal dichalcogenides. Nat. Nanotechnol. 15, 592–597 (2020).

    Article  Google Scholar 

  25. Park, J. Y. et al. A perspective on semiconductor devices based on fluorite-structured ferroelectrics from the materials–device integration perspective. J. Appl. Phys. 128, 240904 (2020).

    Article  Google Scholar 

  26. Peng, Y. et al. Nanocrystal-embedded-insulator (NEI) ferroelectric FETs for negative capacitance device and non-volatile memory applications. Nanoscale Res. Lett. 14, 115 (2019).

    Article  Google Scholar 

  27. Wu, M. & Li, J. Sliding ferroelectricity in 2D van der Waals materials: related physics and future opportunities. Proc. Natl Acad. Sci. USA. 118, e2115703118 (2021).

    Article  Google Scholar 

  28. Li, L. & Wu, M. Binary compound bilayer and multilayer with vertical polarizations: two-dimensional ferroelectrics, multiferroics, and nanogenerators. ACS Nano 11, 6382–6388 (2017).

    Article  Google Scholar 

  29. Scott, J. F. Ferroelectrics go bananas. J. Phys.: Condens. Matter 20, 021001 (2008).

    Google Scholar 

  30. Sarker, B. K. & Khondaker, S. I. Thermionic emission and tunneling at carbon nanotube–organic semiconductor interface. ACS Nano 6, 4993–4999 (2012).

    Article  Google Scholar 

  31. Beebe, J. M., Kim, B., Gadzuk, J. W., Frisbie, C. D. & Kushmerick, J. G. Transition from direct tunneling to field emission in metal-molecule-metal junctions. Phys. Rev. Lett. 97, 026801 (2006).

    Article  Google Scholar 

  32. Wu, M. Two-dimensional van der Waals Ferroelectrics: scientific and technological opportunities. ACS Nano 15, 9229–9237 (2021).

    Article  Google Scholar 

  33. Jeon, H. et al. Hysteresis modulation on Van der Waals-based ferroelectric field-effect transistor by interfacial passivation technique and its application in optic neural networks. Small 16, e2004371 (2020).

    Article  Google Scholar 

  34. Kaushik, N. et al. Reversible hysteresis inversion in MoS2 field effect transistors. npj 2D Mater. Appl. https://doi.org/10.1038/s41699-017-0038-y (2017).

  35. Park, Y., Baac, H. W., Heo, J. & Yoo, G. Thermally activated trap charges responsible for hysteresis in multilayer MoS2 field-effect transistors. Appl. Phys. Lett. https://doi.org/10.1063/1.4942406 (2016).

  36. Wang, H., Wu, Y., Cong, C., Shang, J. & Yu, T. Hysteresis of electronic transport in graphene transistors. ACS Nano 4, 7221–7228 (2010).

    Article  Google Scholar 

  37. Wang, B. et al. Monolayer MoS2 synaptic transistors for high-temperature neuromorphic applications. Nano Lett. 21, 10400–10408 (2021).

    Article  Google Scholar 

  38. Xu, H., Zhou, J., Li, Y., Jaramillo, R. & Li, J. Optomechanical control of stacking patterns of h-BN bilayer. Nano Res. 12, 2634–2639 (2019).

    Article  Google Scholar 

  39. Park, J., Yeu, I. W., Han, G., Hwang, C. S. & Choi, J. H. Ferroelectric switching in bilayer 3R MoS2 via interlayer shear mode driven by nonlinear phononics. Sci. Rep. 9, 14919 (2019).

    Article  Google Scholar 

  40. Guan, Z. et al. Recent progress in two‐dimensional ferroelectric materials. Adv. Electron. Mater. 6, 1900818 (2019).

    Article  Google Scholar 

  41. Qi, L., Ruan, S. & Zeng, Y. J. Review on recent developments in 2D ferroelectrics: theories and applications. Adv. Mater. 33, e2005098 (2021).

    Article  Google Scholar 

  42. Liu, Y. et al. Promises and prospects of two-dimensional transistors. Nature 591, 43–53 (2021).

    Article  Google Scholar 

  43. Li, T. et al. Epitaxial growth of wafer-scale molybdenum disulfide semiconductor single crystals on sapphire. Nat. Nanotechnol. 16, 1201–1207 (2021).

    Article  Google Scholar 

  44. Shen, P. C. et al. Ultralow contact resistance between semimetal and monolayer semiconductors. Nature 593, 211–217 (2021).

    Article  Google Scholar 

  45. Wang, Y. et al. P-type electrical contacts for 2D transition-metal dichalcogenides. Nature 610, 61–66 (2022).

    Article  Google Scholar 

  46. Huang, J. K. et al. High-κ perovskite membranes as insulators for two-dimensional transistors. Nature 605, 262–267 (2022).

    Article  Google Scholar 

  47. Wang, L. et al. Exploring ferroelectric switching in α‐In2Se3 for neuromorphic computing. Adv. Funct. Mater. 30, 2004609 (2020).

    Article  Google Scholar 

  48. Baek, S. et al. Ferroelectric field-effect-transistor integrated with ferroelectrics heterostructure. Adv. Sci. 9, e2200566 (2022).

    Article  Google Scholar 

  49. Wang, S. et al. Two-dimensional ferroelectric channel transistors integrating ultra-fast memory and neural computing. Nat. Commun. 12, 53 (2021).

    Article  Google Scholar 

  50. Kwon, K. C. et al. In-plane ferroelectric tin monosulfide and its application in a ferroelectric analog synaptic device. ACS Nano 14, 7628–7638 (2020).

    Article  Google Scholar 

Download references

Acknowledgements

This work was supported by the National Science and Technology Council (NSTC) of Taiwan through grant 111-2124-M-003-005 (Y.-W.L.), 111-2119-M-008-003-MBK (Y.-W.L.), 110-2634-F-009-027 (C.-L.L.), 110-2112-M-A49-013-MY3 (C.-L.L.) and 110-2112-M-A49-022-MY2 (C.-L.L.). Y.-W.L. thanks the Taiwan Consortium of Emergent Crystalline Materials (TCECM) for its financial support. C.-L.L. thanks the Center for Semiconductor Technology Research (CSTR) for its financial support (the CSTR is from The Featured Areas Research Center Program within the framework of the Higher Education Sprout Project by the Ministry of Education (MOE) in Taiwan). T.H.Y. thanks C. R. Kao at Materials Science and Engineering, National Taiwan University, for his insightful comments on dislocation analyses and the phase transformation model. We thank Y.-M. Chang in The Instrumentation Center at National Tsing Hua University for her skilled TEM operation. We are grateful for technical support on SPEM data acquisition from Y.-C. Kuo, S.-H. Hsieh and C.-H. Chen at National Synchrotron Radiation Research Center (NSRRC) and K.-H. O. Yang at National Taiwan University.

Author information

Authors and Affiliations

Authors

Contributions

T.H.Y. and B.-W.L. contributed equally to this work. T.H.Y. and Y.-W.L. conceived the idea and devised the project. H.-C.H. and Yi-Cheng Chen conducted the material growth. L.Y. and M.W. performed the DFT simulation to deduce the sliding mechanism. T.H.Y., H.-C.H., W.-H.C., H.-C.L. and Y.-C. Chang prepared samples for material characterizations. F.-X.C. and Y.-H.K. performed the scanning tunnelling microscope spectroscopy analysis and the cycling I–V measurement. S.-Z.H. and Yi-Chun Chen conducted PFM investigations and analysed the data. P.-Y.L. conducted the HRTEM and SADP investigation. T.H.Y. did the dislocation analysis and verified the formation and sliding mechanisms of domain boundaries. T.H.Y., H.-C.L. and W.-H.C. designed and fabricated MISM capacitors and FETs. T.H.Y., B.-W.L. and T.-H.K. performed the field-effect measurements. Y.-J.H. supported the FET fabrication and data analysis. T.H.Y., B.-W.L., T.-H.K. and J.-H.C. conducted the P–E measurements, and K.-S.L. examined the validity and interpretation of the results. T.H.Y. and B.-W.L. performed the PUND measurements, and Z.-F.L. and M.-H.L. assisted with the data analysis and interpretation. T.H.Y. and Y.-S.K. did the Wulff net simulation. T.H.Y., H.-C.H., Y.-F.C. and T.-H.L. performed the Raman and PL analyses. T.H.Y. and H.-C.H. conducted the SPEM investigation. H.-C.H. and A.C.C. conducted the AFM analyses. The manuscript was written by T.H.Y., B.-W.L., H.-C.H., K.B.S. and F.-X.C. with detailed discussion with all authors and was revised by C.-L.L., M.W. and Y.-W.L. The overall project was supervised by Y.-W.L.

Corresponding authors

Correspondence to Tilo H. Yang, Chun-Liang Lin or Yann-Wen Lan.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Electronics thanks the anonymous reviewers for their contribution to the peer review of this work.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Controllable synthesis of ST-3R MoS2 epilayers.

a, CVD setup for ST-3R MoS2 synthesis. A two-step CVD synthesis was performed at 800 °C for 3 min and 670 °C for 17 min, respectively. The higher temperature selected for the first step was to intensively vaporize MoO3, which was subsequently brought to the growth substrate by an Ar carrier gas. An additional SiO2/Si substrate was placed in front of the MoO3 powders for gas flow modulation. b, Clarification of the growth conditions for ST-3R MoS2 epilayers by changing temperatures and duration times of the MoO3 zone. The MoS2 flakes produced with different thermal histories, denoted in different color lines, are shown in (c-f). c, ST-3R bilayer MoS2 flakes via a two-step synthesis: 800 °C for 3 min (1st step) and 670 °C for 17 min (2nd step). Scale bar, 20 µm. d, Monolayer MoS2 via a two-step synthesis: 800 °C for 2 min (1st step) and 670 °C for 17 min (2nd step). The Mo amount in this condition is insufficient to promote vertical growth. e, Concentric growth of 3R MoS2 epilayers via a one-step synthesis at 800 °C for 20 min, showing 3R MoS2 flakes commonly observed in CVD synthesis. f, Nucleation of MoS2 does not take place via a one-step synthesis at 670 °C for 20 min. Scale bars in (d-f), 10 µm. The results of (c-f) suggest that the first duration mainly determined the Mo supply for MoS2 growth at 800 °C. However, the growth under the continuous Mo supply at 800 °C led to ordinary concentric-bilayer MoS2. A confined Mo amount and relatively low temperature is a key condition for the ST-3R bilayer MoS2 growth. g, Optical micrographs showing morphological evolutions of ST-3R MoS2 bilayers grown on SiO2/Si substrates with the increasing duration time at 670 °C. The bottom panel illustrates the corresponding growth stages. The epilayer growth can be divided into three stages: (1) formation of six primary dendrites with the preferred orientations along the armchair directions, (2) formation of multiple secondary dendrites based on the primary dendrites, and (3) coalescence of parallelly-oriented dendrites to form a merged epilayer triangle. Scale bars, 20 μm. i, AFM topography image of the marked bilayer region in the optical micrograph (h) of an ST-3R MoS2 flake, showing a monolayer step height of about 0.7 nm for the epitaxial nanoribbons. Scale bar, 10 μm.

Source data

Extended Data Fig. 2 Growth mechanism of ST-3R MoS2 homoepitaxy.

a, Polar plots of the formation energies of MoS2 edges as a function of chiral angle χ (0° ≤ χ ≤ 360°) under a Mo-deficient environment. The interior red triangle represents the final equilibrium shape of a MoS2 epilayer. The 0° corresponds to the armchair direction. b, Schematic of surface and edge diffusion for a MoS3 monomer on 1H-MoS2 surface. The stacking configuration between the 2nd layer and the underlying MoS2 layer is 3R stacking. The enlarged illustration shows possible migration pathways of a MoS3 monomer adsorbed on top of the Mo site, where the black dashed arrows denote the hopping paths for every single hopping event while the red solid arrows are the effective paths of two consecutive hopping events which possess the equivalent vectors parallel to the \(\langle \bar{1}\,2\,\bar{1}\,0\rangle\) (armchair) directions. TMo, the top site of Mo atoms from the 1st layer. H, the hollow site. c, Arrhenius relation of the hopping rates of a MoS3 monomer for the surface and edge diffusion. d, Tuning dimensionality of bilayer nanoribbons by changing the 2nd (tempering) temperature, as denoted by black dashed lines in (c), clearly showing that the bilayer ribbons became thinner with decreasing the temperature. The observed behavior is reasonable since the decrease in growth temperature increased the difference in the diffusion rates between surface diffusion and edge diffusion. The more limited the ability of edge relaxation is, the higher the degree of fractal growth and the thinner the bilayer ribbons turn out.

Extended Data Fig. 3 Shear-transformation-induced 3R polytypic domains and domain boundaries.

a, b, Extended dark-field (top) and bright-field (bottom) TEM images taken with the \((1\,0\,\bar{1}\,0)\) diffraction spot, revealing AB and BA lamellae in the AA-matrix. All the lamellar domains show the preferred orientation along the zigzag directions, which is roughly perpendicular to the growth direction of bilayer nanoribbons. Scale bars, 100 nm. c, Schematic of AB and BA stacking faults forming in the instable AA matrix, where the domain boundaries constitute a pair of Shockley partial dislocations (SPD) with \(\mathop{{b}}\limits^{{\rightharpoonup }}\parallel \langle 10\bar{1}0\rangle\) that is decomposed from a basal dislocation \(\mathop{{b}}\limits^{{\rightharpoonup }}\parallel \langle \bar{1}\,2\,\bar{1}\,0\rangle\). ML refers to monolayer, while BL to bilayer. d, A diffraction pattern from the ST-3R bilayer nanoribbon shown in (e). e, Tilted dark-field TEM image obtained by filtering the \((10\bar{1}0)\) diffraction spot of MoS2, where a BA lamella is embedded in an AA matrix. f, g, Dark-field TEM images from the same region in (e), taken with \({\mathop{{g}}\limits^{{\rightharpoonup }}}_{1}=(2\bar{1}\bar{1}0)\) and \({\mathop{{g}}\limits^{{\rightharpoonup }}}_{2}=(1\,\bar{2}\,1\,0)\) diffraction spots denoted in (d), respectively. Based on the criterion \(\mathop{g}\limits^{\rightharpoonup }\cdot \mathop{b}\limits^{\rightharpoonup }=0\) that is satisfied to make dislocations disappear, the domain boundary \({\mathop{{b}}\limits^{{\rightharpoonup }}}_{1}\) disappears with \({\mathop{{g}}\limits^{{\rightharpoonup }}}_{1}=(2\bar{1}\bar{1}0)\) while the domain boundary \({\mathop{{b}}\limits^{{\rightharpoonup }}}_{2}\) disappears with \({\mathop{g}\limits^{\rightharpoonup }}_{2}=(1\bar{2}10)\). This suggests that BA-AA domain boundaries constitute a pair of Shockley partial dislocations with \(\mathop{b}\limits^{\rightharpoonup }\parallel \langle 10\bar{1}0\rangle\), and the resultant BA domain is a Shockley stacking fault forming in the AA matrix. Scale bars, 20 nm.

Extended Data Fig. 4 Extended STEM-ADF image analysis of shear-transformation-induced 3R domain boundaries.

a, b, STEM-ADF images taken from two neighboring regions in a 3R nanoribbon, showing a BA-AA domain boundary. (M: moiré-like transition zone. PSD: partial screw dislocation). The Burgers vectors of the M zone and the PSD site are parallel to the\(\langle 10\bar{1}0\rangle\) direction. Scale bars, 2 nm. c, Enlarged image showing the transition of atomic-stacking configuration between the AA and BA domains in (b). Scale bar, 1 nm. d, e, STEM-ADF images taken from the regions far from the domain boundary shown in (b). No noticeable lattice misalignment is found in the separated domains. Scale bars, 2 nm.

Extended Data Fig. 5 Diffraction patterns and STEM-ADF images of different 3R stacking domains.

ac, SADPs taken from monolayer, AB-stacking bilayer, and AA-stacking bilayer regions, respectively. The corresponding intensity profiles were extracted along the white dashed square in each SADP image of (a-c), which are consistent with simulated SADPs of 3R stacking polytypes by Kim et al. (Ref. 21). d, e, STEM-ADF images showing atomic arrangements of AB and BA domains, respectively. Scale bars, 2 nm. f, g, Intensity profiles extracted from the red and purple dashed squares in the bilayer and monolayer regions in (d,e).

Extended Data Fig. 6 Evolution of experimental P-E loops of ST-3R MoS2 capacitor and mathematical modeling.

a, P-E loops of the ST-3R MoS2 capacitor at increasing \({{\rm{E}}}_{{{\rm{MoS}}}_{2}}\) ranges, from 0.027 V nm−1 to 0.217 V nm−1. \({{\rm{E}}}_{{{\rm{MoS}}}_{2}}\) is the effective field across the ST-3R MoS2 layer (Supplementary Note 3). b, The experimental and simulated P-E loops, from the corrected Miller et al. model, at \({{\rm{E}}}_{{{\rm{MoS}}}_{2}}\) = 0.109 V nm−1. To consider the asymmetric polarization switching due to the different top and bottom interfacial structures of the ST-3R MoS2 capacitor (top: hBN/MoS2, bottom: graphene/MoS2), the corrected Miller model is used to simulate the P-E loops4,5. c, The experimental data and corrected Miller model estimation of the P-E loops at \({{\rm{E}}}_{{{\rm{MoS}}}_{2}}\) = 0.217 V nm−1. d, Electric-field dependence of 2Ps and 2Pr, extracted from a.

Source data

Extended Data Fig. 7 Temperature-dependent electrical measurements of ST-3R MoS2 FETs.

The SiO2 dielectric thickness for the FETs presented is 300 nm. The applied E fields (defined by applied voltages/the SiO2 thickness) range from 0.033–0.33 V nm−1, which is quantitatively close to that used for the FETs with the 30 nm-thick SiO2 dielectric. a, Transfer characteristics at |VDS|=1 V at increasing VGS sweep ranges, from ±10 V to ±100 V, measured in vacuum and at 298 K. b, Temperature-dependent transfer characteristics at |VDS|=1 V at VGS range of ±100 V. c, The temperature-dependent changes of the Ion/Ioff ratio and the memory window (VMW).

Source data

Extended Data Fig. 8 Concerns of long-term retention arising from charge trapping.

Retention of the program and erase currents of an ST-3R MoS2 FET with floating gate at |VDS|=1 V, for a, an as-fabricated device measured in ambient conditions, b, a vacuum-annealed device measured in ambient conditions, and c, a vacuum-annealed device measured in an N2 atmosphere, respectively. The annealing was conducted at 100 °C for 30 min. Gate pulses of 10 V and −10 V at 1 s were used for the program and erase operations, respectively.

Source data

Extended Data Fig. 9 Statistical analysis on memory performance of ST-3R MoS2 FETs.

a, The dual-sweep IDS-VGS transfer curves repeated for 50 cycles from an ST-3R MoS2 FET device. Cumulative distribution of (b) VMV and (c) the on/off ratio of 50 ST-3R MoS2 FET devices. The bias conditions, |VDS| = 1 V and VGS range = ±10 V, are used to ensure that the FETs reach a maximum VMV and a maximum Ion/Ioff ratio between the program and erase states. The distribution in each graph is fitted with a normal curve. The statistical distribution shows that VMV is around 7 V, and the highest value can reach around 8–9 V. The on-off current ratio between the program and erase states is about 104 and with an average of 6.10 × 103.

Source data

Supplementary information

Supplementary Information

Supplementary Notes 1–6, Figs. 1–14 and Tables 1–3.

Source data

Source Data Fig. 1

Fig. 1a reaction path data.

Source Data Fig. 2

Fig. 2a–c, e–g raw data.

Source Data Fig. 3

Fig. 3b–f, raw data.

Source Data Extended Data Fig. 1

Extended Data Fig. 1h AFM raw data.

Source Data Extended Data Fig. 6

Extended Data Fig. 6a, PE loops; Fig, 6b, PE simulation; Fig. 6c, PE simulation; Fig. 6d, Ps Pr E.

Source Data Extended Data Fig. 7

Extended Data Fig. 7a, IV curves; Fig. 7b IV T curves; Fig. 7c, VMV and on/off ratio; Fig. 7c, analysis raw data.

Source Data Extended Data Fig. 8

Extended Data Fig. 8a, Retention; Fig. 8b, Retention; Fig. 8c, Retention.

Source Data Extended Data Fig. 9

Extended Data Fig. 9a, 50 cycles; Fig. 9b, 50 VMV; Fig. 9c, 50 on/off ratio; Fig. 9c, IV curves.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yang, T.H., Liang, BW., Hu, HC. et al. Ferroelectric transistors based on shear-transformation-mediated rhombohedral-stacked molybdenum disulfide. Nat Electron 7, 29–38 (2024). https://doi.org/10.1038/s41928-023-01073-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41928-023-01073-0

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing