Skip to main content
Nature Communications logoLink to Nature Communications
. 2024 Dec 30;15:10822. doi: 10.1038/s41467-024-55160-7

Dual-role ion dynamics in ferroionic CuInP2S6: revealing the transition from ferroelectric to ionic switching mechanisms

Xingan Jiang 1,#, Xiangping Zhang 2,#, Zunyi Deng 3,#, Jianming Deng 4, Xiaolei Wang 5,, Xueyun Wang 3,, Weiyou Yang 1,
PMCID: PMC11685995  PMID: 39738004

Abstract

Due to its “ferroionic” nature, CuInP2S6 combines switchable ferroelectric polarization with highly mobile Cu ions, allowing for multiple resistance states. Its conductive mechanism involves ferroelectric switching, ion migration, and corresponding intercoupling, which are highly sensitive to external electric field. Distinguishing the dominant contribution of either ferroelectric switching or ion migration to dynamic conductivity remains a challenge and the conductive mechanism is not clear yet. Here, based on polarization switching analyses and first-principles calculations, this work demonstrates that the Cu ion migration pathways enable the formation of a quadruple-well state, determining the conductive mechanism. Accordingly, it favors the manipulation of Cu ion transport in the intralayer and interlayer in a controlled manner, and makes a transition from ferroelectric-dominated to ion-migration-dominated conductivity, by tailoring the electric fields. This work deepens the understanding of ion migration dynamics and conductive switching in ferroionic systems, which is critical for the advancement of memristor-based neuromorphic computing.

Subject terms: Ferroelectrics and multiferroics, Two-dimensional materials


The authors find that the Cu ion migration pathways determine the conductive mechanism of van der Waals ferroelectric CuInP2S6. By tailoring the electric field, they manipulate Cu ions transport within and between layers, revealing the transition from ferroelectric to ionic switching mechanisms.

Introduction

In recent years, there has been a surge of interest in memristor-based neuromorphic computing technologies13. Two-dimensional ferroelectric materials have emerged as highly promising candidates in these fields due to their atomically thin structure and switchable polarization by manipulating the electric field, enabling the energy-efficient information processing and synaptic-like functionalities48.

Among these materials, CuInP2S6 (CIPS) is particularly noteworthy as it exhibits “ferroionic” properties911, that is the combination of switchable ferroelectric polarization with mobile ions. This characteristic allows for the dynamic adjustment of resistance states and self-rectification in memristive devices that are pivotal in memory and logic operations1216. This dual functionality is primarily attributed to the copper (Cu) ions in CIPS, which contribute to both the ferroelectricity characterized by off-center ordering and the ionic conduction, characterized by long-range disordered migration1721. Their close correlation endows this material with novel phenomena, including the existence of quadruple-or sextuple well states2225, polarization reversal against the electric field23,26, anomalous photocurrent reversal27, remarkable ionic conduction12, etc. For example, Brehm et al. 22 revealed that along the direction of the electric field, Cu ions can be migrated from intralayer to a stable interlayer position in the van der Waals (vdW) gap, thereby forming an additional pair of highly polarization states. Neumayer et al. 26 further revealed that the application of electric field can also drive the Cu ions across the vdW gap, which leads to polarization alignment against the electric field. Although these outstanding works have provided a deep understanding of Cu ion migration dynamics and multiple-well states through switching spectroscopy-PFM (SS-PFM) measurements supplemented with density functional theory (DFT) calculations, the relationship between Cu ion migration dynamics and the conductive mechanisms was not addressed, leaving it unclear in this field. Given that Cu ions dynamics induce the ferroelectric switching simultaneously with Cu ions migration in CIPS, the conductive mechanism involves ferroelectric switching, ion migration, and their coupling, all of which are highly sensitive to electric field conditions23,26. It remains a challenge to accurately distinguish whether it is ferroelectric or ion migration that primarily contributes to the dynamic conductivity under varied electric field strengths, directions, and time duration. The concrete evidence linking the dual-role ion dynamics to the conductive mechanisms should be further established, which is the focus of our work.

Recent works have shown that applying an electric field through nanometer tip provides an effective approach to investigate the ionic migration effect and conductive behavior. For polarization switching, Cu ions maintain a relatively ordered position in the lattice18,28,29, while the localized strong electric field under the nanometer conductive tip would lead to the reversible accumulation and depletion of Cu ions under the tip and even complete disordering of Cu ion positions30,31. Another benefit of using nanometer tip is providing a highly asymmetric electrode configuration32, where the conducting behavior is primarily influenced by the nanometer-sized tip and the macroscopic counter-electrode has a minimal impact, greatly reducing the complexity of understanding the mechanisms at play12.

In this work, we used CIPS as the memristor layer and constructed a prototype device with a macroscopic Au bottom electrode and a conductive Pt nanometer tip as the top electrode. Based on the tip-based SS-PFM and first-principles calculations, this work confirms the control of Cu ion transport properties in multiple wells states and reveals their roles in determining the conductive behaviors. It reveals that the intralayer migration leads to ferroelectric-dominated conduction and interlayer migration contributes to ion migration-controlled conduction. The Cu ion migration pathways in the quadruple state can be manipulated in a controlled manner depending on the electric fields, achieving a controlled transition from ferroelectric to ion migration-dominated conductive mechanism. Finally, it also demonstrates that the Cu ion migration can be effectively modulated by the application of mechanical load. Our work provides unique insights into the switching of conductive mechanisms in the ferroionic CIPS.

Results

Ferroelectric-/ion migration-modulated conductive switching behavior

The highly asymmetric electrode configuration was presented in Fig. 1a, where a macroscopic Au was used as bottom electrode and a conducting Pt nanometer tip as the top electrode. The exfoliated CIPS flake was used as memristor layer. The crystal structure can be described as a sulfur framework, with metal cations (Cu and In) and P–P pairs occupying the octahedral voids20, as depicted in Fig. 1b. The unstable Cu ions can occupy three crystallographic sites9, as presented in Fig. 1c: (1) intralayer position Cu1 (upward Cu1u and downward Cu1d), which shifted from the centers of the octahedra, (2) octahedral Cu2 (Cu2u and Cu2d), which located in the centers of the octahedra, and (3) nearly interlayer positions Cu3 (Cu3u and Cu3d), which located in the vdW gap. Under external electric field, Cu ions migrate and can occupy these possible positions, forming multiple wells states. In this work, we aim to investigate the control of Cu ion transport properties in multiple wells states and their roles in determining the conductive mechanisms.

Fig. 1. Ferroelectric-/ion-modulated conductive switching in ferroionic CuInP2S6.

Fig. 1

a The schematic of AFM tip/CIPS/Au on SiO2/Si substrate for electrical measurement. b, c Schematic of layered structured CIPS and possible copper positions labeled Cu1, Cu2, and Cu3. d The switching spectroscopy-PFM (SS-PFM) at ambient environment (~26 °C), with a testing time duration of 10 s. e Ten fast scans of IV curves with a time duration of 1 s for each IV cycle at ambient temperature. The numbers and arrows represent the bias scanning sequence and direction, respectively. The raw data in (e) is adapted from our previous work16 under a Creative Commons licence CC BY 4.0. f The maximum current change with the cycles in two bias direction according to the results of (e). g The IV curves measured at 80 °C with a time duration of 1 s (left panel) and PFM phase at different temperature (right panel).

Figure 1d shows the SS-PFM performed at ambient environment (~26 °C), with a testing time duration of 10 s. Both phase and amplitude reveal hysteresis loops (off-field, Vdc  =  0). A ~180° phase contrast corresponds to the two opposite polarization directions perpendicular to the layer surface. The amplitude signal shows two minima (the red arrow) at the threshold voltage for polarization switching in a double potential wells due to the Cu ions hopping between the intralayer Cu1u and Cu1d positions18. Interestingly, above a threshold voltage of ~+3 V, the amplitude once again shows a trend towards minima (the shaded region), suggesting the onset of quadruple-well due to Cu ion migration across vdW layers18,26. Notably, the activation of quadruple states also significantly influences the conductive behavior. As shown in Fig. 1e and Supplementary Fig. 1, 10 fast scans of full-cycle current–voltage (IV) curves were conducted with a time duration of 1 s for each cycle. Asymmetric conductive switching was observed with the current being larger in negative bias than positive bias due to the asymmetric electrode configurations32. The current in negative bias exhibits a resistive switching window at a threshold voltage of ~−1.5 V and a current peak in positive bias is observed at a threshold voltage of ~+1.5 V due to the dominating ferroelectric switching in a double wells state30. Interestingly, a resistive switching window forms at higher voltages and significantly expands with increased cycles (the shaded region), which identifies the activation of quadruple states due to the enhanced Cu ions migration across the vdW gap. As presented in Fig. 1f and Supplementary Fig. 2, with increased scanning cycles, the current first increases and then starts to decrease after cycle#8 in negative bias direction. Conversely, the current significantly increases after cycle#8 in positive bias direction. Applying a large V, or long t also leads to a rapid current decrease in negative bias direction and a significant current increase in positive bias direction (Supplementary Figs. 35). The significant change in conductive behavior after the turning point under both positive and negative bias serves as proof Cu ion migration across vdW gap and the emergence of the quadruple-well states. In this scenario, ferroelectric polarization charge control of the conductive mechanism is no longer applicable, and ion migration becomes the primary contribution.

The ion migration mechanism was further confirmed by elevating the temperature above the Curie temperature (Tc) to eliminate ferroelectricity. As shown in Fig. 1g and Supplementary Fig. 6, the non-ferroelectric phase can be obtained when the temperature reaches 56 °C, higher than Tc of CIPS. Subsequently, IV curves were measured at higher temperature such as 80 °C, which is sufficient to ensure the non-ferroelectricity. It was strongly confirmed in the non-ferroelectric phase that applying negative bias dramatically decreases the current with low-resistance state (LRS) switching to a high-resistance state (HRS) in a counterclockwise sequence from 1 to 2. Conversely, applying positive bias significantly increases the current with a large resistive switching from HRS to LRS state. This result strongly indicates the ion migration control of conductive behavior. The pronounced accumulation and depletion of Cu ions beneath the nanometer tip12 drastically influences the conduction in two different bias directions in an opposite manner. The current rectification of IV curves can be dynamically manipulated and effectively switched by varying the bias conditions such as voltage magnitude or time duration to activate Cu ion migration (Supplementary Fig. 7).

The abnormal ferroelectric polarization switching behaviors under an electric field

For further investigation on the Cu ion migration dynamics under external electric field, SS-PFM technique was employed33. Ten SS-PFM were performed at a specific location with time duration of 1 s for each cycle (Supplementary Fig. 8). The phase and amplitude at different test cycles are representatively presented in Fig. 2a and b, respectively. For cycle #1, the phase shows the typical hysteresis loops for double wells states. For example, downward polarization (P) can be switched to upward polarization P by a positive electric field (E) at threshold voltage of ~+3 V. P can be switched back to P by applying a negative electric field (E) at threshold voltage of ~−1.5 V. The polarization aligns with the positive E-field (EP or EP) and this behavior is attributed to Cu ions hopping between the intralayer Cu1u and Cu1d positions18,20,34.

Fig. 2. The Cu ion migration dynamics and abnormal switching behaviors under an electric field.

Fig. 2

a, b The phase and amplitude of SS-PFM at different test cycles, respectively. The numbers 1–4 and black arrows represent the bias scanning sequence and direction, respectively. c The calculated potential barrier during Cu ion migration across the vdW layer by nudged elastic band (NEB) method. d The schematic of two stages for the migration Cu ions paths.

Interestingly, for cycle #4, the phase shows the non-typical hysteresis loops. At a threshold voltage of ~+3 V, P was firstly switched to P, and then immediately switched back to P by the E, where the polarization aligns against the electric field (EP), which is the evidence of quadruple-well states due to the Cu ions migration across the vdW gap23,26,35. With further increasing cycles, the threshold voltage for switching from P to P reduces progressively due to the redistributed Cu ions, and down to~1 V for cycle #7. At a threshold voltage of ~+6 V, P can be switched back to P↓ by the E(EP). Under negative E-field E, the polarization also partially aligns against the field (EP) but not as prominently as at positive E-field E. A major difference between them lies in the fact that applying the negative bias firstly promotes the polarization switching (EP) from cycle#1 to cycle#5 and then depresses this process after cycle#5. However, applying positive bias continuously promotes the polarization switching (EP) with the repeated cycles and higher bias voltage. These results are excellently in agreement with the conduction behavior in the positive and negative bias direction, which are solid evidence of the contribution of Cu ions migration across the vdW gap to the conductive mechanism. For cycle #8, polarization switching becomes invisible because the long-range migration across the vdW gap significantly disrupts the ferroelectric order30, which is further confirmed by the formation of topographic bulges up to 30 nm after 10 cycles (Supplementary Fig. 9). Applying a higher Vmax can also induce polarization alignment opposite to the applied electric field (Supplementary Fig. 10).

Figure 2c further shows the migration scenario and the migration barriers during Cu ion migration across the vdW layers. The initial Cu ion is in the central symmetric position of the [S6] cage. The system energy continues to decrease during the process of Cu ion moving to gap and reaches two local minimum energy minimums in the layer 1, which correspond to the Cu1u and Cu2u positions with the P, respectively. The barrier energy reaches the maximum when Cu ions are in vdW gap. Subsequently, Cu ions enter the two local energy minimum in the adjacent vdW layer 2, which corresponds to Cu2d and Cu1d positions with the P, respectively. Overall, the barrier for the intralayer hopping between the Cu1u and Cu2u, or Cu1d and Cu2d is about 0.05 eV. The barrier for the intralayer hopping between the Cu1u and Cu1d positions is about 0.7 eV. The barrier for the interlayer migration between the Cu2u and Cu2d positions is about 1.1 eV. The Cu ion migration pathways can be feasibly manipulated by varying electric field to surpass the migration barrier. As presented in Fig. 2d, two stages can be proposed to illustrate the Cu ions migration pathways under an electric field E (V, t, N), including bias voltage V, time duration t, and repeated cycles N. At stage 1, under low V, short t, or few N, Cu ions can only hop up and down in the intralayer without energy across the vdW gap. At stage 2, with further increased V, t, or N, Cu ions can migrate across the vdW gaps to reach the adjacent layer or the entire layers.

The dynamic conductive behavior under varying electric field

The conductive switching manipulated by the ionic migration dynamics was further confirmed by precise adjustments of electric field E (V, t, N). It is revealed that the intralayer and interlayer Cu ions migration in a quadruple-well state determines the conductive mechanisms. Figure 3a shows the six sinusoidal voltage sweeping cycles varying the bias voltage magnitude and the time taken for six sinusoidal cycles. The corresponding six sets of full-cycle semilog IV curves and the maximum current in two bias directions are presented in Fig. 3b and c, respectively. It has been revealed that the turning point of the conductive mechanism occurs around Vmax = 5 V. Before Vmax = 5 V, the conductive mechanism is primarily dominated by ferroelectric effects due to the intralayer Cu ions hopping. In this range, the current under negative bias is greater than that under positive bias and increases with the increased Vmax. The resistive switching occurs from the HRS to the LRS in a clockwise sequence from 1 to 2. Under positive bias, a distinct current peak can be observed at small positive bias due to ferroelectric switching30. This is especially identified to be the sharp peak of the switching current through the semilog It curves (Supplementary Fig. 11). When Vmax exceeds the turning point of Vmax = 5 V, the conductive mechanism shifts to being dominated by interlayer Cu ion migration across the vdW gap. The current under negative bias begins to decrease sharply, and the resistive switching occurs from the LRS to the HRS in a counterclockwise sequence from 1 to 2. Conversely, under positive bias, the current peak from ferroelectric switching becomes invisible. The current starts to increase significantly and forms a large hysteresis window at a large threshold voltage above 5 V. This result is highly consistent with the Cu ions migration dynamics in a quadruple-well states as confirmed by the polarization switching results in Fig. 2, where at a large threshold voltage of å 5 V, Cu ions start to migrate across the vdW gap or through the entire layers, as evidenced by the alignment of ferroelectric polarization against the electric field. The long-range migration of Cu+ ions can be also verified from the topographic changes where a slightly topographic bump (~2 nm) appears with one semi-cycle (0 V→+5 V→0 V) IV measurement (Supplementary Fig. 12).

Fig. 3. The dynamic transition from ferroelectric to ion migration-dominated conductive behavior under varying electric fields.

Fig. 3

The raw data are adapted from our previous work16 under a Creative Commons licence CC BY 4.0. a The six sinusoidal voltage sweeping cycles varying the bias voltage magnitude and the time taken for six cycles. The time duration is 1 s for each cycle. b The corresponding six full-cycle of semilog IV curves varying bias voltage magnitude. The numbers#1–4 and black arrows represent the bias scanning sequence and direction, respectively. The current change with the six cycles c varying bias voltage and d varying time duration in positive and negative bias direction.

In addition to bias voltage V, the time duration t also significantly influences the Cu ions migration across the vdW gap in a similar manner. Six sets of full-cycle semilog IV curves varying time duration of each cycle are also plotted in Supplementary Fig. 13. The maximum current in negative bias direction is presented in Fig. 3d. With longer time duration, the current first increases and then starts to decrease rapidly in negative bias direction. Conversely, the current significantly increases in positive bias direction. The turning point of the conductive mechanism switching from ferroelectric to ionic migration effect occurs around time duration t = 1.65 s.

The asymmetric Cu ions migration and conductive mechanisms

The conductive switching mechanism under the varied electric field E (V, t, N) can be comprehensively understood by the Cu ion migration dynamics in a quadruple state. At low V, short t, or few N, Cu ions are restricted to hop up or down in the intralayer, unable to migrate across the vdW gap. In this scenario, ferroelectric polarization aligns with the electric field direction, and becomes the dominant conductive switching mechanism. As illustrated in Fig. 4a, when a positive electric field is applied through the bottom Au electrode, the energy band on the Au side shifts downward, where electrons are transported to the Au electrode more difficult due to a higher interface barrier and smaller contact area at the Pt tip/CIPS. Conversely, as shown in Fig. 4b, applying a negative electric field through the top Pt tip electrode shifts downward the energy band on the Pt side, where electrons are easier transported to the Pt tip electrode due to a lower interface barrier and larger contact area at the CIPS/Au side.

Fig. 4. The conductive mechanisms and the asymmetric Cu ions migration across the vdW gap in two opposite bias direction.

Fig. 4

Band diagrams according to the work function of Pt (5.6 eV)50,51, Au (5.1 eV)51, and electron affinity of CIPS (3.7 eV)5 under positive (a) and negative bias (b) with low bias magnitude or short time duration. The schematic of the Cu ions migration across the vdW gap in positive (c) and negative (d) bias direction with high bias magnitude, long time duration or multiple cycles. e The excellent agreement between IV curves and the polarization switching curve due to the Cu ions migration across the vdW gap. The polarization switching curve is taken from that for Cycle #7 in Fig. 2a. The IV curve is the linear plot of semilog IV curve with Vmax = 8 V in Fig. 3b. f The diagram of the accumulation and depletion of Cu ions at interfaces in positive and negative bias direction.

As V, t, or N reaches a certain threshold, Cu ions can migrate across the vdW gaps to adjacent layers or even across the entire layers and this migration process can be evidenced by the ferroelectric polarization aligning against the electric field (EP) or (EP). In this scenario, the interlayer Cu ions migration across the vdW gap becomes the primary contribution to the conductive mechanism. As illustrated in Fig. 4c, d, this migration is significantly more pronounced under positive bias and particularly weak under negative bias. This asymmetry is confirmed by the stronger (EP) switching under positive bias compared to (EP) switching under negative bias in Fig. 2. Figure 4e shows the excellent agreement between conductive behavior and Cu+ migration across the vdW gap. Especially in the positive bias region of the IV curve, Cu ions begin to migrate across the vdW gap at a large threshold voltage, where the current also increases significantly. This result confirms that the extent of this migration determines the conductive behavior. However, distinguishing electronic and ionic conductivity remains challenging due to their strong coupling. Ion migration influences electronic conductivity by forming internal electric fields and modifying Schottky barriers1012. Conversely, electron transport also affects ion migration. When electron transport is hindered due to interfacial barrier, the effective voltage applied inside the bulk decreases, making ion migration more difficult. The time-resolved current on different time scales may be useful to distinguish their contributions. According to the current in response to 10 cycles of 5 V pulse voltage with pulse widths of 0.01 and 0.1 s, a simple estimate suggests that the ionic current is approximately five times greater than the electronic current (Supplementary Fig. 14).

Cu ion transport modulated by the strain engineering

The asymmetric Cu ions migration across the vdW gap under positive and negative biases arises from the asymmetric distribution of Cu ions. Specifically, this involves the local accumulation of Cu ions beneath the tip under positive bias and their depletion under negative bias, as depicted in Fig. 4f. The Cu ions accumulation effectively lowers the contact barrier at the Pt/CIPS side14,15,36 and further facilitates Cu ions migration over long distance. In contrast, the depletion of Cu increases the contact barrier at the Pt/CIPS side, and thus prevents the Cu ions migration over long distance. Therefore, at large V, long t or multiple N, the asymmetric Cu ions migration leads to the opposite conductive manner in two opposite bias directions. Specifically, this means that the current increases significantly under positive bias and decreases sharply under negative bias.

As discussed earlier, the asymmetric Cu ions migration in two opposite bias directions arises from the interfacial contact barrier. Tip force is an effective means of modulating the interfacial barrier through electromechanical coupling effects, including the piezo- and flexo-electric effect due to the non-uniform strain under the tip34,3739. Here, the AFM tip acts as the top electrode and strain source. The IV curves varying time duration are conducted under different tip forces (Supplementary Fig. 15). The linear and semilog IV curves under ~2.8 µN tip force and a time duration of 10 s are representatively shown in Fig. 5a. It is revealed that the tip force significantly increases the current under negative bias to a level that even exceeds the current under positive bias. This effect is more clearly presented in Fig. 5b and Supplementary Fig. 16. At 0.7 µN, as the time duration extends, the current significantly increases under positive bias and rapidly decreases under negative bias due to the accumulation and depletion of Cu ions beneath the tip. This severe depletion of Cu ions leads to lower currents under negative bias at 1.4 µN compared to the currents at 0.7 µN during short time duration. Increasing the tip force significantly prevents the rapid current decrease under negative bias and instead significantly increases the current as time duration extends. The modulation is driven by two key contributions through the synergistic effects of interfacial and bulk effects mechanisms. On one hand, the interfacial barrier can be reduced under non-uniform strain via piezoelectric and flexoelectric effects. As shown in Fig. 5c, under a force of 2.8 µN, the piezoelectric and flexoelectric effects can generate a net downward-pointing strong electric field of up to the MV/cm level. On the other hand, as shown in Fig. 5d, the vdW gap distance can be reduced and the migration of Cu ions into interlayer sites can be facilitated under strain22,40. Overall, the tunable Cu ion migration, responsive to varied electric fields and mechanical loads, highlights its significant potential for multi-state memristor applications.

Fig. 5. Deterministic Cu ion transport via strain engineering.

Fig. 5

a The linear and semilog IV curves with ~2.8 µN tip force and a time duration of 10 s. b The maximum current (Imax) under different tip forces and time duration. c The schematic of modulating Cu ions transport via the piezo- and flexo-electric effect. d The schematic of the reduced vdW gap distance and enhanced ionic migration across the vdW gap under the condition without strain (W/O strain) and with strain (W strain).

Discussion

This work investigates the Cu ion dynamics in a quadruple-well state in ferroionic CIPS and the conductive switching mechanisms. It unveils the Cu ions migration dynamics and their interactions with ferroelectric switching, using advanced tip-based SS-PFM and first-principles calculations. It confirms that the intralayer and interlayer migration of Cu ions leads to the formation of a quadruple-well state and determines the conductive mechanisms. The intralayer migration leads to ferroelectric-dominated conduction and interlayer migration triggers a shift to ion migration-controlled conduction. Through precise control of electric field parameters E(V, t, N)—bias voltage (V), time duration (t), and repeated cycles (N)—this work demonstrates an ability to manipulate the transport properties of Cu ions in the quadruple-well state, achieving a controlled transition from ferroelectric to ion migration-dominated conductive mechanism under varying electric field. Finally, this work also demonstrates the significantly modulated Cu ion migration by the application of mechanical strain. The tunable Cu ion migration, responsive to varied electric fields and mechanical loads, highlights its significant potential for multi-state memristor applications.

Methods

Crystal growth and sample preparation

CIPS single crystals were grown by the chemical vapor transport method20. The stoichiometric elemental precursors were encapsulated in a vacuum-sealed quartz tube (vacuum at the level of ~10−3 Pa). In two-zone tube furnace, the hot zone was kept at ~650 °C and a cold zone at ~600 °C for a reaction time of ~212 h, respectively. For electrical measurement, the as-grown CIPS crystal was mechanically exfoliated and transferred on Au/SiO2/Si substrates. In this work, the thickness of CIPS nanoflakes is ~120 nm.

Switching spectroscopy-PFM (SS-PFM)

The variable-temperature PFM measurements were conducted using a commercial atomic force microscope (Asylum Research MFP-3D) equipped with a temperature controller. The ferroelectric domains were imaged by the Pt/Ir-coated Si cantilever tip driven with an ac voltage (Vac = 2 V) and the drive frequency of ≈350 kHz. The localized hysteresis loops of CIPS were collected using the SS-PFM mode. Pulsed triangular dc driving voltage (Vdc) was used and the dc pulse width was set to 10 ms. A high-frequency ac voltage (Vac = 2 V) was superimposed on the dc voltage. A hysteresis loop consists of 100 pulses as the time duration is 1 s. To void the ambiguity from the electrostatic force for on-field (dc voltage is on) measurements, we focused on the off-field (remanent) data for our analyses. Multiple cycles of hysteresis loops were collected for each measurement at a specific location.

Conductive atomic force microscopy (c-AFM)

In c-AFM measurement, single point IV measurements were performed with ORCA module with compliance current of 20 nA. The Pt/Ir-coated Si cantilever tip was used as the top electrode. The voltage was applied to the conducting Au bottom electrode, which was continuously swept to simultaneously read the current. Multiple cycles of IV measurements were conducted by varying bias voltage magnitude and time duration of each cycle at a specific location.

Piezoelectric and flexoelectric calculation

The strain and strain gradient distributions in CIPS flakes were obtained from theoretical simulations with a tip-force model under a tip force of 2.8 μN41. The Pt/Ir-coated silicon tip was considered as the rigid sphere, tip radius is 25 nm. The Young’s modulus and Poisson’s ratio of CIPS are 25 GPa and −0.06, respectively. For simplicity, the piezoelectric coefficient d33 was set to be −85 pm/V based on previous literature42, and the effective flexoelectric coefficient μeff was set to be 1 nC/m, which is close to the reported values of two-dimensional materials43. The piezoelectric and flexoelectric fields were obtained by using the linear piezoelectric equation and the flexoelectric constitutive equation, respectively.

Density functional theory simulation

To investigate the mechanism of electric field regulation, we employed DFT using the Vienna Ab initio Simulation Package4446. We examined the Cu ion migration barriers between different atomic layers in CIPS and evaluated the minimum energy path using the nudged elastic band (NEB) method47. The effects of various electron exchange–correlation functionals and vdW corrections on the lattice parameters (a, b, c, and β) and lattice volume V are calculated and results are listed in Supplementary Table 1. The Perdew–Burke–Ernzerhof formulation of the generalized gradient approximation48 was chosen to describe the exchange–correlation interaction of electrons, and the DFT-D3 (Becke–Johnson) vdW formulation was used to describe vdW interactions between layers. Wave functions were expanded in a plane-wave basis set with an energy cutoff of 680 eV for structural relaxation and 520 eV for electronic structure calculations. The force convergence criterion for each ion was set to less than 10−3 eV/Å, and a total energy minimization precision of 10−6 eV was adopted. In NEB calculations, one Cu ion was removed to accommodate the migrating ion, with the force convergence criterion tightened to less than 0.04 eV/Å. For the unit cell, we used the 8 × 4 × 4 k-mesh in the Monkhorst–Park scheme in reciprocal space to ensure the convergence for the total energy self-consistent calculations. Valence electrons were defined as follows: Cu: 3p63d104s1, In 4d105s25p1, P 3s23p3, and S 3s23p4. Crystal structures were visualized using VESTA software49.

Supplementary information

Acknowledgements

This work is supported by Ningbo Yongjiang Talent Introduction Programme (2023A-390-G) [X.J.] and Ningbo Top Talent Project (Grant No. 2020-DST-003) [W.Y.], National Natural Science Foundation of China (Grant Nos. 12374080 [X.Y.W.], 92163101 [X.Y.W.], 12474101 [X.L.W.], 12074018 [X.L.W.], 12202056 [X.J.], 12304120 [J.D.]).

Author contributions

X.L.W., X.Y.W., and W.Y. conceived the idea and directed the project. X.J. prepared the samples with the help of J.D. X.J. performed the IV, the PFM measurements, and data analysis. X.Z. performed the analysis of piezo- and flexo-electric field calculation. Z.D. conducted the DFT calculations. X.J., X.Y.W., and W.Y. co-wrote the manuscript. All authors discussed the data and contributed to the manuscript.

Peer review

Peer review information

Nature Communications thanks Wenwu Li and Nikita Ter-Oganessian for their contribution to the peer review of this work. A peer review file is available.

Data availability

The authors declare that all the relevant data within this paper and its Supplementary Information file are available from the corresponding author upon request.

Competing interests

The authors declare no competing interests.

Footnotes

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

These authors contributed equally: Xingan Jiang, Xiangping Zhang, Zunyi Deng.

Contributor Information

Xiaolei Wang, Email: xiaoleiwang@bjut.edu.cn.

Xueyun Wang, Email: xueyun@bit.edu.cn.

Weiyou Yang, Email: weiyouyang@tsinghua.org.cn.

Supplementary information

The online version contains supplementary material available at 10.1038/s41467-024-55160-7.

References

  • 1.Yao, P. et al. Fully hardware-implemented memristor convolutional neural network. Nature577, 641–646 (2020). [DOI] [PubMed] [Google Scholar]
  • 2.Rao, M. et al. Thousands of conductance levels in memristors integrated on CMOS. Nature615, 823–829 (2023). [DOI] [PubMed] [Google Scholar]
  • 3.Huang, Y. et al. Memristor-based hardware accelerators for artificial intelligence. Nat. Rev. Electr. Eng.1, 286–299 (2024). [Google Scholar]
  • 4.Liu, K. et al. An optoelectronic synapse based on α-In2Se3 with controllable temporal dynamics for multimode and multiscale reservoir computing. Nat. Electron.5, 761–773 (2022). [Google Scholar]
  • 5.Wu, J. et al. High tunnelling electroresistance in a ferroelectric van der Waals heterojunction via giant barrier height modulation. Nat. Electron.3, 1–7 (2020). [Google Scholar]
  • 6.Ma, Y. et al. High-performance van der Waals antiferroelectric CuCrP2S6-based memristors. Nat. Commun.14, 1–11 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Wang, Q. et al. Extraordinary tunnel electroresistance in layer-by-layer engineered van der Waals ferroelectric tunnel junctions. Matter5, 4425–4436 (2022). [Google Scholar]
  • 8.Wang, C. et al. Towards two-dimensional van der Waals ferroelectrics. Nat. Mater.22, 542–552 (2023). [DOI] [PubMed] [Google Scholar]
  • 9.Zhou, J. et al. 2D Ferroionics: conductive switching mechanisms and transition boundaries in van der Waals layered material CuInP2S6. Adv. Mater.35, 2370267 (2023). [DOI] [PubMed] [Google Scholar]
  • 10.Sun, Y. et al. Internal ion transport in ionic 2D CuInP2S6 enabling multi-state neuromorphic computing with low operation current. Mater. Today66, 9–16 (2023).
  • 11.Zhong, Z. et al. Robust threshold-switching behavior assisted by Cu migration in a ferroionic CuInP2S6 heterostructure. ACS Nano17, 12563–12572 (2023). [DOI] [PubMed] [Google Scholar]
  • 12.Zhang, D. et al. Anisotropic ion migration and electronic conduction in van der Waals ferroelectric CuInP2S6. Nano Lett.21, 995–1002 (2021). [DOI] [PubMed] [Google Scholar]
  • 13.Liu, Y. et al. Versatile memristor implemented in van der Waals CuInP2S6. Nano Res.16, 10191–10197 (2023). [Google Scholar]
  • 14.Zhu, H. et al. Highly tunable lateral homojunction formed in two-dimensional layered CuInP2S6 via in-plane ionic migration. ACS Nano17, 1239–1246 (2023). [DOI] [PubMed] [Google Scholar]
  • 15.Huang, Y. et al. Cu+ migration and resultant tunable rectification in CuInP2S6. ACS Appl. Electron. Mater.5, 5625–5632 (2023). [Google Scholar]
  • 16.Jiang, X. et al. Manipulation of current rectification in van der Waals ferroionic CuInP2S6. Nat. Commun.13, 574 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Zhou, Y. et al. van der Waals epitaxial growth of one-unit-cell-thick ferroelectric CuCrS2 nanosheets. Nano Lett.24, 2118–2124 (2024). [DOI] [PubMed] [Google Scholar]
  • 18.Zhang, D. et al. Ferroelectric order in van der Waals layered materials. Nat. Rev. Mater.8, 25–40 (2023). [Google Scholar]
  • 19.Zhou, Z. et al. Unconventional polarization fatigue in van der Waals layered ferroelectric ionic conductor CuInP2S6. Nat. Commun.14, 8254 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Liu, F. et al. Room-temperature ferroelectricity in CuInP2S6 ultrathin flakes. Nat. Commun.7, 12357 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Ci, W. et al. Engineering ferroelectric‐/ion‐modulated conductance in 2D vdW CuInP2S6 for non‐volatile digital memory and artificial synapse. Adv. Funct. Mater.34, 2316360 (2024). [Google Scholar]
  • 22.Brehm, J. A. et al. Tunable quadruple-well ferroelectric van der Waals crystals. Nat. Mater.19, 43–48 (2020). [DOI] [PubMed] [Google Scholar]
  • 23.Li, T. et al. Realization of sextuple polarization states and interstate switching in antiferroelectric CuInP2S6. Nat. Commun.15, 2653 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Wang, K. et al. Interface-tuning of ferroelectricity and quadruple-well state in CuInP2S6 via ferroelectric oxide. ACS Nano17, 15787–15795 (2023). [DOI] [PubMed] [Google Scholar]
  • 25.Sivadas, N. et al. Scale-free switching of polarization in the layered ferroelectric material CuInP2S6. Phys. Rev. Res.5, 043074 (2023). [Google Scholar]
  • 26.Neumayer, S. M. et al. Alignment of polarization against an electric field in van der Waals ferroelectrics. Phys. Rev. Appl.13, 064063 (2020). [Google Scholar]
  • 27.Bai, Y. et al. Anomalous photocurrent reversal for the same polarization direction in van der Waals ferroelectric CuInP2S6. PRX Energy3, 023004 (2024). [Google Scholar]
  • 28.Chen, C. et al. Large-scale domain engineering in two-dimensional ferroelectric CuInP2S6 via giant flexoelectric effect. Nano Lett.22, 3275–3282 (2022). [DOI] [PubMed] [Google Scholar]
  • 29.Liu, H. et al. Reversible flexoelectric domain engineering at the nanoscale in van der Waals ferroelectrics. Nat. Commun.15, 4556 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Neumayer, S. M. et al. Ionic control over ferroelectricity in 2D layered van der Waals capacitors. ACS Appl. Mater. Interfaces14, 3018–3026 (2022). [DOI] [PubMed] [Google Scholar]
  • 31.Balke, N. et al. Locally controlled Cu-ion transport in layered ferroelectric CuInP2S6. ACS Appl. Mater. Interfaces10, 27188–27194 (2018). [DOI] [PubMed] [Google Scholar]
  • 32.Yang, C.-H. et al. Electric modulation of conduction in multiferroic Ca-doped BiFeO3 films. Nat. Mater.8, 485–493 (2009). [DOI] [PubMed] [Google Scholar]
  • 33.Jesse, S. et al. Switching spectroscopy piezoresponse force microscopy of ferroelectric materials. Appl. Phys. Lett.88, 062908 (2006). [Google Scholar]
  • 34.Ming, W. et al. Flexoelectric engineering of van der Waals ferroelectric CuInP2S6. Sci. Adv.8, eabq1232 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Xue, F. et al. Emerging van der Waals ferroelectrics: unique properties and novel devices. Appl. Phys. Rev.8, 021316 (2021). [Google Scholar]
  • 36.Xu, X. et al. Multifunctionality of Li2SrNb2O7: memristivity, tunable rectification, ferroelasticity, and ferroelectricity. Adv. Mater.34, 2206022 (2022). [DOI] [PubMed] [Google Scholar]
  • 37.Lu, H. et al. Mechanical writing of ferroelectric polarization. Science336, 59–61 (2012). [DOI] [PubMed] [Google Scholar]
  • 38.Yang, M.-M. et al. Flexo-photovoltaic effect. Science360, 904–907 (2018). [DOI] [PubMed] [Google Scholar]
  • 39.Lun, Y. et al. Ultralow tip-force driven sizable-area domain manipulation through transverse flexoelectricity. Adv. Mater.35, 2302320 (2023). [DOI] [PubMed] [Google Scholar]
  • 40.Yao, X. et al. Anomalous polarization enhancement in a van der Waals ferroelectric material under pressure. Nat. Commun.14, 4301 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Wang, L. et al. Flexoelectronics of centrosymmetric semiconductors. Nat. Nanotechnol.15, 661–667 (2020). [DOI] [PubMed] [Google Scholar]
  • 42.Neumayer, S. M. et al. Giant negative electrostriction and dielectric tunability in a van der Waals layered ferroelectric. Phys. Rev. Mater. 3, 024401 (2019).
  • 43.Yang, Y. et al. Computation of flexoelectric coefficients of a MoS2 monolayer with a model of self-consistently distributed effective charges and dipoles. J. Chem. Phys. 156, 174104 (2022). [DOI] [PubMed]
  • 44.Kresse, G. et al. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci.6, 15–50 (1996). [Google Scholar]
  • 45.Kresse, G. et al. Ab initio molecular dynamics for liquid metals. Phys. Rev. B47, 558–561 (1993). [DOI] [PubMed] [Google Scholar]
  • 46.Kresse, G. et al. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B54, 11169–11186 (1996). [DOI] [PubMed] [Google Scholar]
  • 47.Jónsson, H. et al. Nudged elastic band method for finding minimum energy paths of transitions. in Classical and Quantum Dynamics in Condensed Phase Simulations 385–404 (World Scientific, 1998).
  • 48.Perdew, J. P. et al. Generalized gradient approximation made simple. Phys. Rev. Lett.77, 3865–3868 (1996). [DOI] [PubMed] [Google Scholar]
  • 49.Momma, K. et al. VESTA: a three-dimensional visualization system for electronic and structural analysis. J. Appl. Crystallogr.41, 653–658 (2008). [Google Scholar]
  • 50.Das, S. et al. Enhanced flexoelectricity at reduced dimensions revealed by mechanically tunable quantum tunnelling. Nat. Commun. 10, 537 (2019). [DOI] [PMC free article] [PubMed]
  • 51.Lee, J. et al. Free-standing two-dimensional ferro-ionic memristor. Nat. Commun.15, 5162 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Data Availability Statement

The authors declare that all the relevant data within this paper and its Supplementary Information file are available from the corresponding author upon request.


Articles from Nature Communications are provided here courtesy of Nature Publishing Group

RESOURCES