Introduction

The stable and electrically switchable spontaneous polarization in ferroelectric materials can lead to high-performance non-volatile memories1,2,3,4,5. Ferroelectric tunnel junctions (FTJs) with ultra-thin ferroelectric films have attracted much attention as two-terminal resistive switching devices due to the small cell size, non-destructive readout, and scalability towards large-scale arrays6,7,8,9,10,11. The ferroelectric thin film allows the quantum mechanical tunneling of electrons, with the average tunneling barrier and thus the tunneling electroresistance (TER) modulated by the polarization of the ferroelectric layer12,13,14. A high TER can increase the reliability and decrease the energy consumption in FTJ devices and arrays9,15, and the barrier height or width modulation has been proven to improve the TER value16. By using a top metal electrode with a high work function, an increased average barrier height is obtained17, which leads to a TER of about 100. Recently, barrier height modulation above 1 eV is demonstrated in Cr/CuIn2S6 (~4 nm)/graphene FTJs by tuning the graphene layer thickness18, leading to a TER over 107. Besides, it is also effective to modulate the effective barrier width to further enhance the TER. The niobium-doped strontium titanate (Nb:SrTiO3, NSTO) has been used as a semiconducting bottom electrode to provide an extra Schottky-like barrier in the high-resistance state, owing to the depletion region at the ferroelectric/semiconductor interface19, and the TER over 106 is achieved in a Pt/BaTiO3 (BTO)/NSTO device20.

Besides a high TER, the scaling of ferroelectric layer thickness is also critical for FTJs because it can reduce the program voltage and latency. However, as FTJs continue to scale down towards atomic scale, it is quite challenging to maintain a large TER when the ferroelectric layer thickness is down to the one-nanometer range, because the TER is limited by the structural instability of the ferroelectric phase and the enhanced depolarization field at nanometer-scale thickness21,22. For a 1-nm-BTO-based FTJ which is achieved using a large compressive strain imposed by a NdGaO3 substrate6, the TER is only about 2. FTJs based on 1-nm Hf0.8Zr0.2O2 (HZO) with strain confinement show enhanced ferroelectricity23, but the TER is still only about 200, far from that of commercial flash memories.

Recently, the strong out-of-plane ferroelectricity is demonstrated in layered bismuth oxide with samarium (Sm) substitution24. The samarium bonds stabilize the ferroelectricity even when the film is down to 1 nm, which shows strong potential for high-performance non-volatile memories. In this article, we demonstrate metal/ferroelectric/semiconductor FTJs based on Sm-substituted bismuth oxide (BSO), which can achieve a TER over 7 × 105 with a 1-nm ferroelectric BSO film. Such TER is more than three orders of magnitude higher than the TER values of previously reported FTJs with 1-nm ferroelectric layers. We also show that the TER can be further enhanced as the thickness of the BSO layer increases. The TER over 109 has been obtained in FTJs with 4.6-nm-thick BSO films, which is among the highest TER reported in thin-film FTJs at room temperature. The giant TER is attributed to both ferroelectric barrier height modulation and the barrier height/width modulation at the surface depletion region of NSTO19, thanks to the large polarization maintained at small thickness for the BSO layer24. These BSO-based FTJs can further support multi-level cells and analog memories. By varying the reset stop voltages, 32 distinct resistance states (5 bits of data storage) are realized in a single FTJ. Moreover, the devices demonstrate linear conductance modulation, more than 5 × 109 endurance cycles, and over 10 years of retention time. These BSO-based FTJs hold high promise towards reliable, high-performance, and low-power non-volatile memory and in-memory computing applications.

Results

Structure and ferroelectric properties of ultra-thin BSO films

Layered Sm-substituted bismuth oxide (Bi2–xSmxO3) is a recently discovered material with ferroelectric states driven by a lone pair of electrons, which can maintain a large polarization value down to 1 nm24. The large polarization at such a small thickness (t) is beneficial for high-performance FTJs. We have grown (001) epitaxial Bi1.8Sm0.2O3 (BSO) thin films on NSTO substrates by a sol-gel method, followed by lithographically patterning the top electrodes (see Methods Section for details). The schematic diagram of the Au (60 nm)/Cr (15 nm)/BSO (1 nm)/NSTO device structure is shown in Fig. 1a, b. The atomic model of the BSO thin film is shown in Supplementary Fig. 1, and the atomic arrangement of the BSO film with thickness down to ~1 nm is confirmed by the high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) (Fig. 1c). Due to the stress of the substrate and surface, the lattice of the ultrathin BSO film along the c axis is stretched to a certain extent, resulting in a stable polarization25,26. At the same time, X-ray reflectivity measurements and the fitting results confirm that the thickness of the BSO film is ~1 nm with a surface roughness of 0.173 nm24, which is consistent with the theoretical model and the TEM image.

Fig. 1: Structure of the Au/Cr/BSO/NSTO-based FTJs and ferroelectric properties of the 1-nm BSO films.
figure 1

a Schematic illustration of the FTJ structure. b Atomic-level diagram of the Au (60 nm)/Cr (15 nm)/BSO (1 nm)/NSTO FTJ structure. While the Sm atoms are distributed randomly in the structure, we provide one possible BSO structure obtained from a previous study24. c The HAADF-STEM image of a BSO film grown on NSTO substrate, with the BSO thickness of ~1 nm. d The polarization hysteresis loop (left axis) and switching current (right axis) of the 1-nm BSO, obtained by the PUND-mode measurement under an electric field at 625 Hz frequency. e The out-of-plane phase of the PFM measurement, shown in color scale after a box-in-box writing with a tip bias in the 1-nm-thick BSO film. The entire detection area is 8 μm × 8 μm. The tip biasing is noted in the corresponding area, and the region without the note is unbiased. f Off-field local PFM measurements for BSO films with the thickness of 1 nm, including both phase (left axis) and amplitude (right axis) hysteresis loops.

In the positive-up-negative-down (PUND) measurement mode (Supplementary Fig. 2), the FTJ with 1-nm-thick BSO layer shows a standard macroscopic hysteresis loop, by integrating the switching current curve (Fig. 1d). The remanent polarization up to 16.6 μC/cm2 has been obtained, which is the foundation for achieving nonvolatile switching with a high TER. Piezoresponse force microscopy (PFM) measurements also demonstrate clear ferroelectric properties of the 1-nm-thick BSO films, with the out-of-plane phase showing obvious ferroelectric domain walls (Fig. 1e). The phase of the unbiased region is consistent with that of the region biased at +7 V, which proves that the ultrathin BSO films have intrinsic ferroelectric properties. The amplitude vs. voltage curve of the off-field local PFM presents a butterfly shape (Fig. 1f), which is a typical piezoelectric characteristic24. For BSO films with different thicknesses, the coercive fields and remanent polarizations are summarized in Supplementary Table 1.

Giant TER of BSO-based FTJs

We leverage the high polarization of the BSO layer maintained at 1 nm thickness to achieve a high TER in FTJs. We employ the metal/ferroelectric/semiconductor device structure, which allows the modulation of both the height and width of the barrier19. The semiconductor substrate is STO with 0.7 wt% Nb doping, which provides low resistivity for decreasing the low-resistance state (LRS) resistance, and allows the efficient modulation of surface potential and the formation of the depletion region using the applied voltage or the polarization in the ferroelectric layer, as reported previously20. We first perform measurements using quasi-static voltage sweeps, by applying voltages on the metal side with the NSTO substrate grounded. These BSO-based FTJs show repeatable nonvolatile switching between the low-resistance state and high-resistance state (HRS) for multiple cycles (Fig. 2 and Supplementary Fig. 3). In FTJs with 1-nm BSO layer, we obtain a TER up to 7 × 105 by using 2 V set voltage and −4.5 V reset voltage (Fig. 2a, b). The resistance shows gradual change with set and reset voltages without abrupt switching, and the switching characteristics are highly repeatable (inset of Fig. 2a). We also find that a higher reset voltage than set voltage is needed, because of the extra depletion region formed at the BSO/NSTO interface during the reset process, which will divide a portion of the total voltage, as reported previously13,19. Furthermore, for FTJs with 4.6-nm-thick BSO, the TER is 108 with −6 V reset volage (Supplementary Fig. 4), and reaches up to 1.2 × 109 with −7 V reset voltage (Fig. 2c, d). The compliance current of 1 mA is applied in the set process to prevent the large tunneling current from affecting the device performance. The thickness effects on the HRS and LRS values are summarized in Fig. 2e, with the data of the FTJ with 2.1-nm BSO extracted from Supplementary Fig. 5. We observe that there is no significant change in the LRS (around 10 kΩ), while the HRS resistances increase with BSO thickness (from about 1.3 GΩ to 3.1 TΩ), which results in a higher TER with thicker BSO layers.

Fig. 2: Giant TER values in BSO-based FTJs.
figure 2

a Quasi-static I–V sweeps for multiple cycles of a FTJ with BSO thickness of 1 nm. The voltages are applied on the metal side with the NSTO substrate grounded. Inset: cumulative probability distribution of LRS and HRS. b Over 7 × 105 TER achieved with −0.1 V read voltage. c Quasi-static I–V sweeps for multiple cycles of a FTJ with BSO thickness of 4.6 nm. Inset: cumulative probability distribution of LRS and HRS. d Over 1.2 × 109 TER achieved with −0.1 V read voltage. e Thickness effect on the TER values of BSO-based FTJs. The data of the FTJ with 2.1-nm BSO is extracted from Supplementary Fig. 5. The colored ranges in the box chart (25–75%) show the interquartile range (IQR), which is between the first and third quartile of the data distribution, and the error bars show the range within 1.5 IQR. f Comparison of TER values among ultra-thin FTJs with different ferroelectric materials and thicknesses. The purple squares are results of BTO-based devices from refs. 6,20,27. The magenta diamonds are results of PZT-based devices from refs. 29,30. The green triangle is result of the CuInP2S6 device from ref. 18. The blue hexagons are results of HZO-based devices from refs. 23,32. The brown triangles are results of BFO-based devices from refs. 28,31, and the black triangle is the result of the α-In2Se3-based device from ref. 33. For FTJs with 1-nm-thick ferroelectric layer, our BSO-based FTJs show three orders of magnitude improvement in TER. Our FTJs with 4.6-nm-thick BSO show the highest TER of 1.2 × 109.

We benchmark the room-temperature TER values of our FTJs against those based on other types of ultrathin ferroelectric layers, including ABO-type perovskites (where A is a rare-earth or alkaline-earth metal, B is a transition metal, and O is oxygen)6,20,27,28,29,30,31, binary oxides23,32, and van der Waals layered materials18,33 (Fig. 2f). Intriguingly, the high TER (7 × 105) of the FTJs based on 1-nm BSO film is more than three orders of magnitude higher than previously reported TERs of FTJs based on other 1-nm-thick ferroelectric materials: the TER is only ~2 for FTJs based on 1-nm BTO6, and is ~200 for FTJs based on 1-nm HZO23. While the metal/ferroelectric layer/semiconductor structure has been previously demonstrated19,31,34, such high TER has not been reported for FTJs based on 1-nm-thick ferroelectric films. When the thickness of the ferroelectric layer is down to 1 nm, the “dead layer” in the film, which has low dielectric constant can decrease the polarization21. In our device, Sm can help overcome the ferroelectric structural instability and depolarization field24, and maintain a high polarization of 16.6 μC/cm2 in BSO at 1 nm thickness. The high polarization of BSO is beneficial for modulating the barrier height and width, leading to such high TER values. We further demonstrate the highest TER up to 1.2 × 109 for FTJs based on the 4.6-nm-thick BSO film, which is more than 20 times higher than the state-of-the-art BTO-based FTJs (~5 × 107)27 and CuInP2S6-based FTJs (~107)18. Such high TER values will decrease the leakage current and increase the reliability of the FTJs.

Analysis of giant TER for BSO-based FTJs

Analysis of the FTJ switching mechanism is demonstrated in Fig. 3. Upon contact with zero biasing, the different work functions between Cr (~4.5 eV) and the NSTO substrate will result in charge transfer, leading to a depletion region in NSTO and a built-in potential (Vbi) at the surface19. The energy band bending and energy barrier also changes with the polarization of the ferroelectric layer20, and the effect is further enhanced with a thin BSO layer. When the BSO polarization is pointing from Cr/Au to NSTO, the positive bond charges in BSO will drive the electrons in NSTO to accumulate at the surface. The large polarization of BSO can eliminate the surface depletion in NSTO, and reduce the energy barrier (Fig. 3a). In this case, the FTJ is in LRS, and the structure is similar to metal/ferroelectric/metal FTJs, which supports large tunneling currents, with the I–V characteristics well fitted by the direct tunneling (DT) model (Supplementary Fig. 6)35. When the polarization is pointing from NSTO to Cr/Au, the negative bond charges in BSO will enhance the depletion near the surface of the NSTO. In this case, the average ferroelectric barrier height increases. Meanwhile, an additional energy barrier at the ferroelectric/semiconductor (F/S) interface due to the surface depletion becomes more prominent in the conduction especially when the ferroelectric film is ultra-thin (Fig. 3b)20. The device is then in the HRS state, with the I–V characteristics showing strong asymmetric and rectifying characteristics due to the F/S energy barrier (Fig. 2b, d). The ideality factors (n) are extracted as 2.1 and 1.93 for FTJs with 1-nm-thick and 4.6-nm-thick BSO, respectively (Fig. 3c), based on the junction transport model through an energy barrier (Supplementary Note 1). The deviation of n from unity (ideal thermionic emission over a Schottky barrier) can be attributed to the thermally assisted tunneling in HRS, which has been observed in other NSTO-based Schottky-like junctions36. The large polarization of the BSO film effectively enhances both the average ferroelectric barrier height and the depletion region width, which highly suppress the HRS current, leading to larger TER values.

Fig. 3: Origin of giant TER.
figure 3

a, b The energy band diagrams of FTJs in (a) LRS and (b) HRS states. In (a), the polarization is pointing to the degenerately doped NSTO side, and the positive bond charges in BSO lead to electron accumulation at the surface of NSTO. In this case, the electrons only need to tunnel through the ultra-thin ferroelectric barrier, and the FTJ is in LRS. In (b), the polarization is pointing to the metal side, and the negative bond charges in BSO will deplete the electrons at the surface of NSTO. In this case, an additional energy barrier at the F/S interface increases the effective tunnel barrier width, and thus the device is in HRS. c Fitting to the measured I–V curves for FTJs in HRS with 1-nm and 4.6-nm BSO layers, from which the ideality factors are extracted as n = 2.1 and n = 1.93 for FTJs with 1-nm-thick and 4.6-nm-thick BSO, respectively, suggesting the existence of an extra energy barrier and thermally assisted tunneling mechanism in HRS. d The projected density of states onto the Ti 3d orbitals for the interface layers from DFT+U calculation. When the polarization is pointing to the NSTO side (downward arrow), all the interface (IF) layers are metallic. When the polarization is pointing to the Cr/Au side (upward arrow), some interface layers of NSTO convert into insulating due to the carrier depletion. IF-1 denotes the most neighboring NSTO layer adjacent to the ferroelectric layer.

To confirm the electronic structure at the interface, we perform density functional theory plus Hubbard U (DFT + U) calculations for the heterostructure model composed of one-layer bismuth oxide and four-layer NSTO (see Methods section). The interface model can be found in Supplementary Fig. 7. Because the Ti 3d states are the dominating states near the Fermi level in doped SrTiO337,38, we show the projected density of states (DOS) onto the Ti 3d orbitals for the interface layers in Fig. 3d. In the case where the ferroelectric polarization is pointing toward the NSTO, all the interface layers are metallic. However, as the ferroelectric polarization is switched, some layers at the interface convert into insulating due to the carrier depletion, which implies that the width of the tunneling barrier will be increased according to previous reports39. Our DFT + U calculations also confirm the elimination of energy barrier in LRS and the existence of extra barrier in HRS, consistent with the energy band diagrams shown in Fig. 3a, b. We further demonstrate that the effect of Sm atoms on the interfacial DOS and electronic properties at the interface is negligible (Supplementary Fig. 8). In addition, we have simulated the interface between bismuth oxide and Cr. The charge density difference in Supplementary Fig. 9 shows that the charge transfer can be modulated by the ferroelectric switching, which indicates that the Cr/BSO interface could also contribute to the large TER.

Multi-level cells and analog memory properties of BSO-based FTJs

The giant TER combined with low variation can lead to multi-level cells towards high-density non-volatile memories on chip. By applying alternating reset and set pulses and measuring the resistances after the pulses, we demonstrate 20 cycles of switching in the FTJs with 1-nm-thick BSO ferroelectric layer (Fig. 4a). In addition, by fixing the set pulses at +2 V, and gradually increasing the reset voltage pulse amplitudes from −2 V to −4.5 V, the LRS of ~ 10 kΩ, and the HRS from 200 kΩ to 2 GΩ have been demonstrated. Then by fine-tuning the reset voltage from −1.5 V to −4.5 V with a step of −0.1 V, 32 distinct resistance states (5 bits of data storage) are achieved without any write-verification or write-termination processes (Fig. 4b). For each reset voltage, multiple cycles of resistive switching have been performed, and cycle-to-cycle variation has been summarized. While some resistance overlap exists between certain neighboring resistance states when considering the worst-case cycle-to-cycle variation, such small variation is tolerable for many in-memory computing applications40. When even higher accuracy of multi-level storage is necessary, reducing the resistance levels to 16 (4 bits of memory) can avoid resistance overlap even for the worst-case cycle-to-cycle variation. Other types of non-volatile memories such as resistive random-access memories can also realize similar resistance levels, but the closed-loop programming technique is used due to device variation and nonlinear conductance tuning41. When even more resistance levels are desirable in our FTJs, write-verify or write-termination techniques can also be used to further reduce the variation, although there could be additional costs in circuit area, power, and delay42,43.

Fig. 4: Multi-level cell and analog properties of FTJs based on 1-nm-thick BSO.
figure 4

a Different resistance states achieved by applying reset pulses with different amplitudes. By applying alternating reset and set pulses and measuring the resistances after the pulses, we demonstrate repeatable resistive switching between LRS and different HRS states. b Cumulative probability distribution of 32 distinct resistance states achieved by applying increasing reset pulse amplitudes from −1.5 V to −4.5 V with a step of −0.1 V, obtained from multiple cycles of measurements for each reset pulse amplitude. c The conductance change with increasing pulse numbers, using identical pulses of +1.4 V for potentiation and −1.6 V for depression. The pulse duration is td = 60 ns. d Improved potentiation and depression process using ISPP, showing repeatable and linear conductance modulation. The potentiation process uses a series of pulses with voltages increasing from +1.2 V to +1.95 V, with a step of +0.05 V, and the depression process uses pulses from −1.4 V to −2.9 V, with a step of −0.1 V. e Resistance change with pulse duration using +2 V set pulses and −4.5 V reset pulses, showing that the TER increases with the pulse duration, and TER over 105 can be achieved using 50 μs pulses. f Resistance change with voltage measured using the ISPP scheme, showing a larger resistance tuning range when gradually increasing the peak pulse amplitude (VP).

We further explore linear tuning of multiple resistance states in the FTJs by gradually switching the ferroelectric domains9,40. We first measure the conductance states of BSO-based FTJs with increasing numbers of identical pulses, and find that the conductance is continuously modulated with +1.4 V pulses for potentiation and −1.6 V pulses for depression (Fig. 4c). The nonlinearity factor α is then extracted as 0.234 for potentiation and −0.494 for depression, where fully linear conductance modulation should have α of 0 (Supplementary Note 3). To further improve the linearity, we then apply another programming scheme: incremental step pulse programming (ISPP). The potentiation process uses a series of pulses with voltages increasing from +1.2 V to +1.95 V, with a step of +0.05 V, and the depression process uses pulses from −1.4 V to −2.9 V, with a step of −0.1 V (Fig. 4d). We observe repeatable and linear conductance modulation, with α extracted to be −0.106 for potentiation, and −0.195 for depression. The linear potentiation and depression are demonstrated in multiple FTJs using the ISPP programming scheme (Supplementary Fig. 10). The consistent linear conductance modulation in these FTJs is important not only for high-density memory, but also for in-memory computing and neuromorphic computing9,29. While the ISPP programming technique is more suitable for linear conductance modulation, the identical pulse technique can save the additional control circuits and energy consumption. Therefore, it could be preferred for resource-constrained systems.

The resistance states are also modulated by the voltage pulse duration. For a typical 1-nm-BSO-based FTJ, we find that the TER increases with a longer pulse duration (Fig. 4e). With +2 V set voltage and −4.5 V reset voltage, the device achieves TER over 105 using 50 μs pulses, similar to the programming time for previously reported FTJs with similar device sizes18,31,44. We show that the TER is inversely proportional to the switching speed, and thus the operating speed can be highly improved if the device operates with a relatively small TER. As such, our device provides high flexibility of operation by trading off the TER and switching speed. It is also possible to use metal electrodes with lower work functions, or use smaller electrodes to further enhance the switching speed40,45. Using the ISPP programming technique, the resistance tuning range can be significantly enlarged by increasing the peak pulse amplitude (VP) due to the stronger ferroelectric polarization (Fig. 4f and Supplementary Fig. 11).

Endurance, retention, and device-to-device variation of BSO-based FTJs

Switching endurance, data retention, and device-to-device variation are critical for FTJs towards frequent set/reset, long-term stability, and large-scale integration40. We measure the endurance properties by first applying high programming voltage pulses of 2.5 V for set and −5 V for reset, and show TER over 105 for about 5 × 104 endurance cycles (Supplementary Fig. 12). To achieve even higher endurance, we slightly decrease the programming voltages to 2.2 V for set and −3.5 V for reset, and demonstrate endurance up to 109 cycles with no significant degradation of TER (TER over 300). In addition, the total switching endurance up to 5 × 109 with TER ~ 100 for the 1-nm-BSO-based FTJ is demonstrated (Fig. 5a). While the TER degrades after 5 × 109 cycles, the I–V sweeps measured during the endurance switching cycles prove that the device is still functional (Supplementary Fig. 13), and the resistance window can be recovered by slightly increasing the voltage amplitude (Fig. 5b). The degraded TER could be attributed to the “dead layer” formed at the interface due to oxygen vacancy aggregation during the repetitively switching, which will limit the domain switching34. Furthermore, we measure the endurance of another FTJ based on 1-nm BSO, using 2.5 V set pulses and −5 V reset pulses until the device degrades (Supplementary Fig. 14). We find that the TER remains above 2 × 105 for over 105 cycles, and then starts to degrade. Nevertheless, the total endurance cycle over 5 × 109 cycles is still demonstrated. The higher endurance of BSO-based FTJs compared with that of previous reports (Supplementary Table 2) could be related to the high-quality and stable BSO film with less grain boundaries24,46.

Fig. 5: Endurance, retention, and device-to-device variation of FTJs based on 1-nm-thick BSO.
figure 5

a Measured endurance properties showing more than 5 × 109 cycles under +2.2 V/−3.5 V programming pulses. b HRS and LRS resistances after 5 × 109 programming cycles, showing that the device can maintain TER over 100 by applying slightly larger programming voltages. c Measured retention property showing that TER over 105 can be maintained for more than 10 years. d HRS, LRS, and TER measured for 35 devices, showing small device-to-device variation.

The retention properties are measured for more than 10 hours for both LRS and HRS (Fig. 5c), which translates to over 105 TER maintained for more than 10 years. We also measure the retention properties for several intermediate states (Supplementary Fig. 15), showing consistent and stable resistance states. The long retention of these BSO-based FTJs could be attributed to the Sm-induced stable tetragonal-like phase structure24, which minimizes the degradation of polarization due to the depolarization field as in previous ultra-thin FTJs46,47. We further measure the HRS, LRS, and TER of 35 FTJs with 1-nm BSO layer (Fig. 5d). High uniformity among different devices with TER over 105 has been demonstrated, which is important towards large-scale array integration10.

Discussion

The NSTO substrate has been a common substrate for constructing high-performance FTJs6,19,45. While the depletion region at the surface of NSTO may lead to a larger resistance compared with a metal contact, the high switching speed of 300 ps and low programming energy below 10 fJ have been demonstrated in FTJs using the NSTO substrate40. Furthermore, although currently the NSTO substrate instead of Si substrate is used, it is promising to obtain BSO films on substrates compatible with complementary–metal–oxide–semiconductor (CMOS) fabrication processes. For example, ultrathin SrTiO3 or similar films can be grown on silicon-on-insulator (SOI) wafers using molecular beam epitaxy (MBE)48,49, or bonded to the Si substrate through the ion slicing, or the “smart-cut” process50,51, which could also be used for BSO film growth. Furthermore, by using a water-soluble Sr3Al2O6 as the sacrificial layer, the SrTiO3 and BiFeO3 thin films have been released from the growth substrate in water, and transferred to a Si substrate. Such a technique may also be used to transfer the BSO films. Besides, we demonstrate the initial results for the growth of the BSO film on metal-on-oxidized-Si substrates such as the Au/SiO2/Si substrate (Supplementary Fig. 16). These results show high potential for the integration of BSO films onto Si substrates, and for on-chip memories using BSO-based FTJs.

In oxide-based materials such as BSO, the filamentary switching due to vacancy motion should also be considered. However, in these BSO-based FTJs, we do not observe the initial forming process that is required for most of the filamentary switching memories without special forming-free designs52,53. We also measure the temperature-dependent I–V characteristics in both LRS and HRS (Supplementary Fig. 17), which shows that the current increases with temperature in both HRS and LRS, with a larger current increase in HRS. Such trend in LRS is contrary to the filamentary switching which should have a decreasing current with higher temperature due to the metallic conduction through the conductive filament54. Through fitting, we find that direct tunneling is dominant in LRS, and thermally assisted tunneling is dominant in HRS, with the energy barrier of 0.1 eV extracted for LRS, and 1.25 eV extracted for HRS, consistent with ferroelectric switching (Supplementary Note 1.2). Furthermore, we measure the capacitances of the FTJs in LRS and HRS (Supplementary Fig. 18). The lower capacitance in HRS than that in LRS, and the capacitance–voltage relationship that fits well to the depletion capacitance model suggest the existence of a built-in potential and a depletion region (Supplementary Note 2). Therefore, ferroelectric switching should be the dominant mechanism in these BSO-based FTJs.

In summary, we demonstrate that BSO-based FTJs with down to one-nanometer ferroelectric layer can maintain a large TER over 7 × 105, which is at least three orders of magnitude higher than the TER of other FTJs with 1-nm-thick ferroelectric barrier. The FTJs with 2.1-nm and 4.6-nm BSO layers achieve TER values over 108 and 109, respectively, which are higher than commercial flash memories. The DFT + U calculations confirm that the effective modulation of energy barrier owing to the large polarization of the BSO layer is essential to achieve the high TER. These BSO-based FTJs also show desirable analog memory properties by exhibiting 32 resistance levels and high linearity in conductance tuning. Furthermore, they show high endurance cycle, long retention, and small device-to-device variation, making them highly promising for low-power, high-reliability, and high-density non-volatile memories and in-memory computing.

Methods

Material growth and device fabrication

BSO films were prepared by chemical solution method. Bi(NO3)3·5H2O and Sm(NO3)3·6H2O were dissolved in glycol methyl ether solvent at the molar ratio of 10:1 in turn, thoroughly mixed and left for 4 h. The global bottom electrode is the conductive substate NSTO (0.7 wt% Nb-doped SiTiO3) with a thickness of 500 μm. The precursor solution was dropped onto the substrate and spun for 30 s at the speed of 5000 rpm. Then annealing was performed at 90, 270, and 600 °C, with the annealing time of 10 min, 5 min, and 30 min, respectively. On the epitaxially grown BSO film, the top circular 15 nm Cr/60 nm Au electrodes with diameters of 10 μm were lithographically defined.

Material characterization

The TEM images of the BSO-based FTJs were obtained by scanning the specimens using the Titan G2 60–300 instrument, where the specimens were prepared by a focused ion beam tool (FEI Helios NanoLab 600i) operated at the voltage of 2–30 kV. The polarization−electric field (P − E) loops were measured using the ferroelectric measurement system (aixACCT TF-Analyzer 3000) at the frequency of 0.01 − 1 MHz. Atomic force microscopy machines from both Asylum Research (MFP3D) and Bruker (Dimension Icon) were used to obtain the piezoresponse force microscopy images. The box-in-box patterns were written on the BSO film by applying the voltage of +7 V or −7 V to the corresponding regions, respectively. The X-ray diffraction (XRD) pattern of the BSO film was collected by an X-ray diffractometer (PW 3040-X’Pert Pro, PANalytical, Cu Kα, λ = 1.54 Å).

Electrical characterization

Electrical measurements of the FTJs were performed using a Keithley 4200A-SCS semiconductor parameter analyzer (for I − V measurements) and Agilent B1500A semiconductor parameter analyzer (for C − V measurements) connected to a probe station under atmospheric pressure and at room temperature. For the BSO-based FTJs, source measurement units (SMUs) with preamplifiers were used in quasi-static electrical measurements, pulse measurement units (PMUs) with remote preamplifier (RPM)/switch modules were used for the pulsed measurements, and the multi-frequency capacitance measurement units (MFCMUs) were used for C − V measurements. During the measurements, the two probes were connected to the top electrode on BSO film and the NSTO substrate under BSO layer, respectively.

First-principles calculations

The VASP code55,56,57, incorporated with Perdew-Burke-Ernzerhof (PBE) functionals and a cutoff of 450 eV, was employed to carry out the non-spin polarized density functional theory calculations. To study the electronic structure at the interface between the ferroelectric layer and NSTO, we employed a supercell structure composed one-unit cell of ferroelectric bismuth oxide and four layers of NSTO. To simulate the epitaxial growth of the bismuth oxide on NSTO, the in-plane lattice constant a was constrained to the optimized lattice constant of NSTO (3.978 angstrom) from our first-principles calculations. We used the rotationally invariant DFT + U method58 to take into account correlation effects on 3d orbitals of transition metal ions. The commonly used U = 5.0 eV and J = 0.64 eV were applied to Ti and Nb 3d orbitals59. In the self-consistent field and DOS calculations, the k-grids of 8 × 4 × 1 and 12 × 6 × 2 were used, respectively. Since the concentration of Nb in the actual NSTO was very low (0.7% in weight percent) and cannot be well simulated by standard first-principles calculations, we replaced one Ti atom out of eight by one Nb atom to model the Nb doped SrTiO3. Specifically, one Ti atom at the interface was replaced with one Nb atom.