Junghyeon
Hwang
a,
Hunbeom
Shin
a,
Chaeheon
Kim
a,
Jinho
Ahn
*b and
Sanghun
Jeon
*a
aSchool of Electrical Engineering, Korea Advanced Institute of Science and Technology, 291 Daehakro, Yuseong-gu, Daejeon 34141, Republic of Korea. E-mail: jeonsh@kaist.ac.kr
bDivision of Materials Science and Engineering, Hanyang University, Seoul, Korea. E-mail: jhahn@hanyang.ac.kr
First published on 2nd November 2024
Ferroelectric materials, characterized by their polarization switching capabilities, have emerged as promising candidates for non-volatile memory applications due to their fast operation speeds, low switching energies, and remarkable scalability. Among these, hafnia-based ferroelectrics are particularly noted for their compatibility with complementary metal-oxide-semiconductor (CMOS) technology. However, the development of high-quality ferroelectricity in ultra-thin films, essential for low-voltage operations and high-density integrations, remains challenging. This study introduces a novel low-damage metallization process designed to fabricate ultra-thin (sub-5 nm) ferroelectric films exhibiting exceptional ferroelectric properties and reliability. The process, compatible with standard CMOS techniques, achieves a significant remnant polarization (Pr) of 40 µC cm−2 and low leakage currents, alongside enhanced retention characteristics. Crucially, it substantially mitigates the wake-up effect, often attributed to oxygen vacancy redistribution at the interface. Through comprehensive analyses utilizing electron energy loss spectroscopy (EELS), geometric phase analysis (GPA) and X-ray photoelectron spectroscopy (XPS), we demonstrate that our process effectively reduces oxygen vacancies and dislocations at the top interface of the ferroelectric film. The enhanced reliability of ferroelectric random-access memory (FeRAM), evidenced by improved sensing margins and consistency in ferroelectric properties, marks a substantial improvement over the conventional method. To precisely measure reliability characteristics, we propose a new retention model that considers charge screening over time. Moreover, circuit-level simulations via non-volatile memory simulator (NVSim) validate the process's integration potential with existing CMOS technologies, affirming its suitability for advanced, high-density memory configurations without compromising performance or energy efficiency. The findings from this study pave the way for broader applications of nanoscale high-quality dielectric thin films, extending beyond ferroelectric materials to various technological domains requiring advanced material solutions.
The integration of ferroelectric memories with CMOS technology—including devices such as ferroelectric field effect transistors (FeFETs),8–14 ferroelectric tunnel junctions (FTJs),15–20 and ferroelectric random-access memory (FeRAM)21–24—has been propelled by advances in hafnia-based ferroelectrics.25 Specifically, FeRAM, which substitutes the DRAM capacitor with a hafnia-based ferroelectric, utilizes readout circuits to detect switching currents triggered by changes in ferroelectric film polarization, a few nanometers in thickness. Its non-volatile nature significantly cuts down the refresh time needed, positioning hafnia-based FeRAM as a potential successor to DRAM by addressing its inherent limitations.26
However, the application of hafnia-based ferroelectrics in ultra-thin films required for low-voltage operation (2 V) and high-density integration faces substantial hurdles due to high coercive fields (>1 MV cm−1) necessitating thickness scaling below 5 nm.27 Although hafnia ferroelectric thin films as thin as 1 nm have been reported, these have only been demonstrated on the silicon substrate, which limits their practical application in FeRAM.19 Moreover, when scaled to less than 10 nm, hafnia ferroelectrics exhibit accelerated degradation of ferroelectric properties, complicating the achievement of sufficient sensing margins and increasing the depolarization field, thereby challenging the construction of reliable devices with ultra-thin ferroelectric films.28,29
To overcome these obstacles, recent innovations have been made in various technological facets, including electrode engineering,28,30 plasma treatment,31 interfacial layer engineering,32,33 and various annealing processes.34–37 Particularly, X. Lyu et al. reported ferroelectricity and anti-ferroelectricity in 4 nm and 4.5 nm thick HfZrO films, respectively.38 They used WN electrodes to stabilize ferroelectricity through interfacial engineering. At 4 nm thick HfZrO with Hf:
Zr = 1
:
1, a remnant polarization (Pr) value of 13.8 µC cm−2 was achieved and 4.5 nm thick HfZrO with Hf
:
Zr = 1
:
3, a saturated polarization (Ps) value of 28.1 µC cm−2 was achieved. However, the leakage current density of 4 nm-thick film was over 10−1 A cm−2, which is quite a large value to the minimum requirement for the DRAM capacitor technology (10−7 A cm−2 at 0.8 V). Also, the reliability characteristics and cycling effects of the films were not reported and the origin of the good ferroelectricity was not well understood. S. Oh et al. reported the 2Pr value of 42.5 µC cm−2 at 5 MV cm−1 field with 2.5 nm thick HfZrO film and Pt as a bottom electrode.39 They asserted that the orthorhombic phase could not occur because the dead layer increased leakage current and that Pt can successfully reduce the formation of the dead layer. Pt is not, however, compatible with the CMOS process, making it unsuitable for use in ferroelectrics based on hafnia, where CMOS compatibility is the main benefit. The majority of the research, including the ones mentioned above, places a strong emphasis on the Pr value and provides little attention to the study of reliability features.
Recently, Cheng et al. discovered that in the annular dark-field mode, a sophisticated imaging technique, using Cs-corrected scanning transmission electron microscopy, significant dead layers are formed at the boundary between the top electrode and HZO layer.40 The dead layer was mainly composed of the t-phase. Meanwhile, the formation energy of the t-phase is known to be lower than that of the o-phase and m-phase when the oxygen vacancy concentration is high, stabilizing the t-phase in environments with high oxygen vacancy concentrations.41,42 They proposed that a damaging effect occurs during the deposition of TiN layers, resulting in the formation of an t-phase interfacial layer.
In this article, we provide a low-damage top electrode metallization method that can fabricate ultra-thin (5 nm) ferroelectric films with superior reliability and ferroelectricity. The low-damage top electrode metallization method was implemented using an atomic layer deposition (ALD) process characterized by low electron temperatures and high plasma density, as measured by optical emission spectroscopy (OES). With an ultra-thin ferroelectric capacitor, we achieved a 2Pr value (40 µC cm−2), low leakage current, and excellent retention properties by a straightforward and CMOS-compatible technique. Also, we discovered a significant reduction in the wake-up effect brought on by the redistribution of defects at the interface. It shows that the low-damage metallization process can reduce the defects of the thin film by preventing damage to the ferroelectric film during top electrode deposition. The dislocation of the ferroelectric film was visualized with a high-resolution cross-sectional transmission electron microscopy (HR-TEM) with electron energy loss spectroscopy analysis (EELS) and geometric phase analysis (GPA). To accurately assess the effects of the interfacial layer on retention, we proposed a new retention model that incorporates the charge screening effect. This model has allowed us to simulate and analyze the short-time retention characteristics of ultra-thin ferroelectric films comprehensively. Ultimately, this study examines the feasibility of deploying ferroelectric capacitors fabricated through this innovative process in FeRAM applications.
Fig. 1(a) displays the OES for Ar/NH3 plasma as a function of the NH3 gas flow rate, indicating the dissociation degree of NH plasma through the ratio of NH to Ar peaks. Enhanced dissociation of NH3 plasma facilitates the generation of ionized ammonium ions and free nitrogen atoms, essential for the activation of reactants and effective film deposition in the PEALD process. As shown in Fig. 1(b), increasing the NH3 flow enhances the density of NH3 molecules in the plasma, boosting the likelihood of molecular collisions and thus promoting dissociation. However, this also increases the probability of recombination reactions, reducing dissociation levels beyond an NH3 flow of 250 sccm. Fig. 1(c) demonstrates that the electron temperature, which is proportional to the emission ratio of Ar 750 nm (2p1–1s2) to Ar 812 nm (2p9–1s5), can be effectively measured. In this analysis, the specific emission wavelengths of 750 nm and 811 nm for argon were chosen for their predominant excitation mechanisms: the former primarily through collisions exciting argon from the ground state, and the latter via excitation of metastable argon atoms.46 These were utilized to derive measurements of electron temperature.
Fig. 1(d) illustrates that the intensity of the Ar 750 nm emission line, proportional to plasma density, reveals an inverse relationship between electron temperature and plasma density.47 High plasma density reduces the mean free path of electrons, increasing collision frequency and energy loss at higher densities, consequently reducing ion bombardment energy and minimizing film damage. We have compared devices fabricated under low damage conditions (150 W, 250 sccm) with those under conventional conditions (300 W, 250 sccm), demonstrating significant differences in electron temperature and plasma density.
Ferroelectric capacitors are essential for next-generation ferroelectric random-access memory (FeRAM) technologies, which require high polarization, minimal leakage currents, and the absence of ‘wake-up’ effects for enhanced reliability and power efficiency. To this end, we performed polarization–voltage (P–V) and current–voltage (I–V) measurements with bipolar cycling. The P–V characteristics of the metal–ferroelectric–metal (MFM) capacitors employing both conventional and low-damage metallization processes are depicted in Fig. 2(a). These capacitors were characterized by a driving voltage of 1.75 V and a frequency of 100 kHz. Initial measurements with the conventional metallization process displayed antiferroelectric-like behavior, transitioning to ferroelectric characteristics after the wake-up process. This behavior is detrimental to FeRAM applications due to the necessity of a wake-up procedure to ensure adequate sensing margins.
Contrastingly, the low-damage metallization processed device exhibited a superior initial remnant polarization (2Pr) of 40 µC cm−2, with minimal wake-up effects, thereby enhancing its suitability for FeRAM use. This approach not only preserved the high polarization values but also significantly reduced the extent of wake-up, a crucial factor for device stability and reliability. As illustrated in Fig. 2(b), the Pr in devices employing the conventional process increased by 30% relative to the initial value, whereas devices processed with the low-damage technique exhibited a threefold increase in Pr value.
Multiple mechanisms, including phase transitions, domain wall depinning, and ferroelastic switching, have been proposed to explain the wake-up effect in hafnia ferroelectric films.48 These effects are largely attributed to the redistribution of oxygen vacancies at the interface, which are typically considered defects and can elevate the leakage current. The comparative analysis of leakage currents, as shown in Fig. 2(c), demonstrated a reduction by at least two orders of magnitude in the low-damage process compared to the conventional method, particularly in the negative bias region. This significant reduction is likely influenced by top electrode engineering, given the identical conditions of the bottom electrode and thermal processing across the devices.
To investigate the interface properties of the device, we employed pulse switching measurements, a method important for the analysis of ferroelectric interfaces as underscored in preceding investigations.49 The measurements began with the application of a pulse of substantial magnitude (4 MV cm−1) directed to completely invert the ferroelectric polarization. This was followed by the application of pulses with varied magnitudes (ranging from 3 MV cm−1 to 4 MV cm−1) in the reverse direction to discern the ferroelectric interface properties through the switching current induced by these pulses, as per the polarization reversal theory.50 The switching current, Isw(t), as a function of time, t, is governed by the equation:
![]() | (1) |
![]() | (2) |
Fig. 3(c) presents the relationship between I0sw and Ea, from which the Ci is deduced by evaluating the slope of the graph and x-axis intercept. When employing both a negative read pulse (pink) and a positive read pulse (purple), the low-damage processed device consistently demonstrated higher Ci values. (Fig. 3(d)) The pronounced slope associated with the polarization switching current suggests an enhancement in the interfacial capacitance for capacitors fabricated using the low-damage process, implying a comparatively thinner interfacial layer than that formed by conventional metallization. This interfacial layer, often a result of non-stoichiometric compound formation, becomes critically impactful with diminishing ferroelectric thickness, potentially leading to retention degradation and increased operational voltage.
Furthermore, pulse-switching characteristics were analyzed across wake-up cycles to ascertain the Ci value. Utilization of the low-damage process exhibited a stable Ci value up to 107 cycles, contrasting with the significant wake-up behavior seen in the conventional process (ESI,† Fig. S1). This observation aligns with the trend of increasing polarization with cycling as shown in Fig. 2, suggesting that defects within the interfacial layer, particularly oxygen vacancies, migrate or diffuse under field cycling, thereby restructuring the built-in field. The increase in Ci value is attributed to domain depinning or phase-transition-induced reduction of the dead layer. Beyond 107 cycles, a decrease in Ci values was observed in both samples, attributed to further defect generation. Milan Pešić et al. reported through simulations that field cycling in a TiN/HfO2/TiN stack MFM capacitor induces accelerated oxygen vacancy formation at the TiOx interface towards the electrode. These oxygen vacancy defects increase non-stoichiometric films at the interface, thereby diminishing the Ci value.
Unlike single-crystal ferroelectrics, which typically conform to the Kolmogorov–Avrami–Ishibashi (KAI) model,51 the switching behavior in polycrystalline hafnia-based ferroelectrics can be effectively modeled using the nucleation limited switching (NLS) approach.52,53 This model enables the analysis of domain sizes and defects within the ferroelectric layer.
Fig. 4(a) and (b) illustrate the ΔP(t)/2Ps values for MFM capacitors processed through conventional and low-damage processes, respectively. These graphs plot the change in polarization as a function of both the magnitude and width of the applied switching pulses, with detailed methodologies described in ESI,† Fig. S2. The device utilizing the low-damage process exhibits a pronounced, steep slope in the switching time versus the normalized polarization graph, suggesting a localized, concentrated switching event. Conversely, the device processed using conventional techniques shows a gentler slope, indicating a more dispersed polarization switching across the ferroelectric layer. This dispersion correlates with the slopes observed near the coercive field (Ec) in the P–E curve depicted in Fig. 1(a), further implying that multiple domains are simultaneously switched in the low-damage processed device compared to the conventional processed device.
The fitting of polarization data to the NLS model is governed by the following equations:
![]() | (3) |
![]() | (4) |
The extracted parameter t1, displayed in Fig. 4(e), reveals an increased value in the device employing the low-damage process, suggesting a reduction in the disorder-induced local fields that typically promote early polarization switching under weaker electric fields.54 These local fields, arising from defects such as oxygen vacancies, interfacial imperfections, and domain boundaries, are mitigated in the low-damage process.
Furthermore, the parameter w from the distribution, depicted in Fig. 4(f), provides additional information about the domain size distribution within the films. A smaller w value denotes a narrower distribution of domain switching times, which correlates with larger domain sizes. Devices utilizing the conventional process exhibited larger w values, indicating smaller domain sizes and more restricted domain-wall propagation due to the aforementioned microstructural disorders. Conversely, the low-damage process, by minimizing these defects, supports the growth of larger and more uniformly distributed domains across the ferroelectric film. These findings underscore the critical role of film processing techniques in determining the microstructural and electrical properties of hafnia-based ferroelectric devices. By optimizing the processing conditions to reduce microstructural disorders, the low-damage process not only enhances the domain size but also promotes a more homogeneous domain distribution, crucial for the development of high-performance ferroelectric devices.
To investigate the leakage current characteristics attributed to defects, we conducted measurements of the current density–voltage (J–E) characteristics for MFM capacitors fabricated using both conventional and low-damage processes. These measurements spanned temperatures from 298 K to 358 K, in increments of 15 K, as shown in Fig. 5(a) and (b). In both sample sets, an increase in leakage current was observed with rising bias voltage and temperature, indicative of field-enhanced thermal emission of electrons. Remarkably, devices employing the low-damage process exhibited leakage currents approximately two orders of magnitude lower across all tested temperatures compared to their conventional counterparts.
The leakage mechanism was analyzed using the Poole–Frenkel emission model, which posits that thermally excited electrons from a metal layer transit into the dielectric without encountering significant energy barriers at the interface and are subsequently moderated by traps within the dielectric. The leakage current density, governed by this model, is expressed as:
![]() | (5) |
Fig. 5(c) and (d) display the Poole–Frenkel plots for the MFM capacitors processed via conventional and low-damage techniques, respectively. The characteristic linear dependency of log(J/E) on E confirms the Poole–Frenkel emission of the leakage current, described by:
![]() | (6) |
![]() | (7) |
![]() | (8) |
As depicted in Fig. 5(e) and (f), b(T) is plotted as a function of q/kT. Utilizing eqn (8), the barrier height Φt was derived from the slope of the b(T) versus q/kT plot. The fitting results indicated Φt values of 0.39 eV for devices processed using the conventional method and 0.64 eV for those using the low-damage process. The disparity in trap depths is attributed to the different defect profiles in these films, including oxygen vacancies, grain boundaries, and the presence of carbon, hydrogen, and nitrogen atoms, which can form traps. According to the Poole–Frenkel model, which favors the predominance of the shallowest trap, the low-damage process effectively minimizes the formation of such shallow traps, manifesting only deeper traps.
Furthermore, to elucidate the impact of oxygen vacancies on the electrochemical properties of HZO thin films, electron energy loss spectroscopy (EELS) spectral imaging was conducted, specifically targeting the O K edge. Fig. 6(a) and (b) display the EELS spectra from the low-damage and conventional samples, respectively, captured vertically from the top to the bottom of the films as depicted in Fig. 6(c) and (d). The O K energy-loss near-edge structure (ELNES) in HZO is characterized by a double-peak profile, labeled a (around ∼534 eV) and b (∼538 eV), with peak a's intensity and sharpness being particularly sensitive to the crystallinity and oxygen vacancy content within the HfO2 matrix. Given that both sets of HZO films were subjected to identical thermal conditions, any deviations observed in the O K ELNES are primarily attributed to differences in oxygen vacancy concentrations or defects induced during the deposition of the top TiN layer.
To quantitatively assess the presence of oxygen vacancies within these films, the relative intensities of peak a within the O K ELNES were analyzed using a nonlinear least squares (NLLS) method to calculate the b/a ratio across the film's profile.55 This b/a ratio acts as an indicator of oxygen vacancy content, where higher ratios indicate increased vacancy presence due to the relative reduction in peak a's intensity. Remarkably, the analysis revealed a significant decrease in the content of oxygen vacancies or defects in samples processed with minimized damage. Unlike the uniform distribution observed in conventional samples, a spatially non-uniform distribution of oxygen vacancies was identified in the low-damage samples, with vacancies preferentially aligning along the [00−1] crystallographic direction and exhibiting reduced concentrations toward the film's surface. These observations are corroborated by X-ray Photoelectron Spectroscopy (XPS) results, further substantiating the spatial variation in oxygen vacancy distribution within these engineered thin films (ESI,† Fig. S4). In addition, we employed scanning transmission electron microscopy (STEM) and geometric phase analysis (GPA) to compare the microstructural quality of ferroelectric films produced by low-damage and conventional metallization processes, observing that the low-damage process results in less strain and fewer dislocations (ESI,† Fig. S5).
![]() | (9) |
This depolarization field, Edep, originates from incomplete charge screening, primarily due to the finite screening length within the metal electrodes58 and the presence of an interface layer.59 In scenarios where complete screening is achieved, the internal field within the ferroelectric material is nullified. For a capacitor comprised solely of ferroelectric material, electrode charges fully compensate the polarization, resulting in no residual field within the ferroelectric under zero bias. Consequently, increased polarization screening correlates with reduced depolarization fields and improved retention characteristics.
To assess the impact of the depolarization field on retention loss, the device's retention was evaluated over a timeframe extending from 20 nanoseconds to 10 seconds. Initially, devices were subjected to a wake-up process at room temperature, involving 1000 cycles of electric field application at a frequency of 100 kHz. Subsequently, a pulse of −1.75 V amplitude and 10 µs width was employed to induce unidirectional switching of the ferroelectric material, followed by the application of a pulse of 1.75 V amplitude and 10 µs width after intervals ranging from 10 ns to 10 s. The current response to the second pulse comprises both the switching and non-switching components of the ferroelectric material, whereas the response to the third pulse contains only the non-switching component. By subtracting the current induced by the third pulse from that induced by the second pulse, we isolate and quantify the contribution from the remnant polarization. This analytical method allows for a precise assessment of the ferroelectric material's retention characteristics by focusing solely on the remnant polarization, effectively filtering out noise and other non-relevant electrical responses.
Our investigation reveals that devices employing a low-damage process retain over 90% of initial polarization after 10 seconds, in contrast to conventional process devices, which maintain only 40% after the same duration. (Fig. 7(a)) To elucidate these observations, we refined a previous retention model based on the premise that depolarization-induced back-switching occurs on a domain-by-domain basis within the ferroelectric layer, which is conceptualized as being divisible into a multitude of regions (N0, with N0 ≫ 1). The depolarization field opposes the polarization in the ferroelectric layer, leading to back-switching of domains and a reduction in net polarization over time. To model the retention behavior, we consider that the ferroelectric capacitor can be divided into N0, each capable of independent back-switching due to the depolarization field. The back-switching of these regions over time causes a gradual decrease in the total polarization P(t). The model suggests that the reverse switching of these regions through nucleation and growth of reverse domains is an independent process for each region, causing a gradual decrease in residual polarization and consequently in the depolarization field as modeled by Lou's feedback equation.60
We start with the fully polarized state at t = t0 (immediately after the removal of the external poling field). The probability that a single region has not back-switched after a time interval Δtn is given by:
![]() | (10) |
![]() | (11) |
![]() | (12) |
![]() | (13) |
As each region back-switches, the overall polarization decreases, and thus the depolarization field reduces according to eqn (9). This introduces a feedback mechanism into the system, as the depolarization field depends on P(t), which changes over time due to back-switching.
Previous models, which overlooked the effects of screening polarization, invariably overestimated the depolarization field and consequently exaggerated the predicted retention loss. To rectify this, our updated model incorporates the dynamics of charge injection following polarization reversal, which initiates screening of the polarization at the interface layer. This screening, triggered by either the depolarization field or an internal field, underscores that insufficient initial charge injection leads to a pronounced depolarization field. Accordingly, the duration of charge injection significantly influences the magnitude of the depolarization field, thereby affecting retention positively. The model has been visually conceptualized through an equivalent circuit, wherein Ri denotes the resistance encountered by charge injection at the interface, and Cf signifies the capacitance of the ferroelectric element. (Fig. 7(b)) This formulation enables a depiction of charge screening via a simple RC model, leading to a revised equation for the depolarization field that accounts for charge screening effects as follows.
![]() | (14) |
The screening charge builds up at the interface, reducing the effective depolarization field. By introducing this charge screening effect, we’ve adjusted the depolarization field equation to reflect a more accurate depiction of retention behavior.
![]() | (15) |
By incorporating charge screening, we account for the reduction of the depolarization field over time, which slows down the back-switching process and improves retention.
The adjusted formula shows a moderated depolarization field over time due to the screening charge, leading to a slower rate of polarization decrease (Fig. 7(c)). The detailed parameters utilized for the fitting of our model are delineated in Fig. 7(d), highlighting the derived values of intrinsic switching time (t∞), activation field (α), and time constant (τ). Remarkably, τ exhibited a significant discrepancy between devices, being calculated at 9 µs for those subjected to a low-damage process, as opposed to 700 µs for devices manufactured via conventional methodologies. This difference is attributed primarily to the resistance at the interface layer (Ri), given that the ferroelectric capacitor (Cf) remains consistent across both device types due to the uniform thickness and material used. The marked variance in τ thus suggests a pronounced reduction in the interfacial layer for low-damage processed devices, inferred from considerations of tunneling resistance, which is dependent on the thickness and material of the barrier. This reduction in the interfacial layer is critical, as it directly contributes to the superior short-term retention characteristics observed in ultra-thin ferroelectric films processed with the low-damage technique. Such an approach effectively mitigates the growth of the interfacial layer, which is pivotal in preserving the ferroelectric properties necessary for reliable FeRAM operation. Furthermore, an exhaustive examination of long-term retention characteristics corroborates the enhanced performance of the latter, as detailed in ESI,† Fig. S6.
To ascertain the efficacy of the low-damage metallization process for FeRAM technology, we investigated its impact on the sensing margin, a critical performance parameter. Fig. 8(a) and (b) outline the pulse scheme utilized for stress application and subsequent readout. This scheme involved a stress pulse to polarize the ferroelectric material, followed by a readout pulse to assess the polarization state. The memory window, crucial for state-of-the-art embedded DRAM (eDRAM) technology, is defined by the charge differential between the read bit 1 and read bit 0 pulses, which should ideally exceed 50 fC µm−2 as per industry benchmarks.
Fig. 8(c) and (d) illustrate the endurance cycling and retention characteristics of the ferroelectric capacitors, key indicators of device reliability and performance under repeated use. By applying bipolar pulses with a magnitude of 3.5 MV cm−1 and a duration of 10 µs, we evaluated the memory window across numerous cycles. Both devices demonstrated endurance exceeding 1010 cycles, as shown in Fig. 8(c). Retention tests were conducted at 358 K and involved 107 bipolar cycles to wake up the devices before measurement. The memory window for the device employing the conventional process was observed to exceed 200 fC µm−2. Conversely, the memory window for the same device under a short retention time displayed less than 100 fC µm−2 due to substantial polarization switching triggered by a strong depolarization field, indicative of a low-quality surface state as detailed in Fig. 7. In contrast, the device processed using the low-damage method exhibited superior non-volatility, maintaining its performance over a projected 10-year lifespan, attributable to an enhanced memory window and robust retention characteristics.
To thoroughly assess the performance of the memory window in both conventional and low-damage treated devices, we employed a cumulative distribution function (CDF) analysis. Fig. 8(e) presents the memory window distribution for conventional devices, while Fig. 8(f) shows the corresponding distribution for low-damage treated devices. For this statistical examination, we calculated the mean and standard deviation of the memory window across a large number of measured devices (n = 34). The memory window represents the difference in charge storage between the high and low polarization states. The lower percentiles and upper percentiles (depicted by the black line) were used to highlight the spread of the data across different devices. In both figures, the memory window data are plotted as a function of CDF, with the mean and standard deviation values explicitly displayed. The results show a mean memory window of 82.7 fC µm−2 with a standard deviation of 3.9 fC µm−2 for conventional devices and a significantly improved mean of 342 fC µm−2 with a reduced standard deviation of 3.2 fC µm−2 for low-damage devices. The lower and upper percentile lines closely follow the reference line, indicating that the data are well-distributed without significant outliers. The smaller standard deviation in the low-damage treated devices suggests improved uniformity in the memory window across devices, which can be attributed to the more controlled and less damaging processing environment. This tighter distribution is essential for applications that require consistent performance, as it reduces the risk of variability that could compromise device reliability.
To assess the electrical characteristics of FeRAM at the circuit level, we utilized non-volatile memory simulator (NVSim), a tool renowned for its efficacy in simulating non-volatile memory systems.61 This simulation aimed to compare the performance of MFM capacitors fabricated via conventional and low-damage processes.
Fig. 9(a) illustrates the schematic of the FeRAM read-out operation, highlighting the charge sharing mechanism between the bit line and the MFM capacitor, critical for distinguishing the memory states. As depicted in Fig. 9(b), the memory states, C0 and C1, represent data 0 and data 1, respectively. The C0 state is characterized solely by the Ps value, whereas the C1 state comprises both the 2Pr and the Ps value. The sensitivity of the FeRAM read-out is quantified by the bit line sensing voltage (ΔVBL), calculated as follows:
![]() | (16) |
Firstly, the low-damage process was shown to effectively increase the sensing margin and memory window of FeRAM devices, crucial metrics for their application in high-density memory arrays. Our experimental results demonstrated that capacitors fabricated with this process maintain high polarization values and exhibit reduced wake-up effects, thereby ensuring stable and reliable memory performance over extended periods. This is a marked improvement over capacitors manufactured using conventional processes, where poorer interface quality and higher defect densities lead to performance degradation.
Furthermore, through advanced characterization techniques such as X-ray diffraction, electron energy loss spectroscopy, and high-resolution transmission electron microscopy, we have identified that the superior performance of low-damage process devices is largely due to the minimized formation of detrimental interfacial layers and defect structures. This fine control over the material properties at the nanoscale ensures that the essential characteristics of ferroelectricity are preserved, even under the demanding conditions imposed by modern electronic applications.
The integration of these ferroelectric capacitors with existing CMOS technologies, as explored through simulations with non-volatile memory simulator (NVSim), indicates that the low-damage process not only supports the scalability of ferroelectric memories but also aligns with the industry's push towards higher density memory arrays. Even in high-density configurations, the detectable bit line sensing voltage further validates the utility of the process in enhancing device performance without compromising energy efficiency or operational voltage requirements.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4tc02460e |
This journal is © The Royal Society of Chemistry 2025 |