NaNbO3 is regarded as an ideal model for elucidating lattice distortion in dielectrics due to its complex structural phase transition under various external fields. In particular, it has garnered significant attention because of its irreversible electric-field-induced transition from antiferroelectric to ferroelectric phases and dielectric relaxation with varying temperatures. At present, numerous studies have explored the evolution of NbO6 octahedra in NaNbO3, while investigations into the role of Na+ remain scarce during phase transition. Moreover, traditional microstructure characterization techniques, such as x-ray diffraction, generally fail to capture the displacement details of lighter Na+ ions, making it particularly challenging to analyze their influence on the phase transition of NaNbO3. In this work, Na1+xNbO3 series with different Na contents are constructed by adjusting the nonstoichiometric ratio of A-site Na atoms. Based on in situ Raman spectra with varying electric fields and temperatures, an in-depth insight into the phase transition is realized, accompanied by revealing the related physical mechanism. The findings will make contribution to identify the regulation of Na+ on the phase transition in NaNbO3 from the perspective of phonon evolution, further clarifying the conversion origin among various electrical orders. In addition, it provides a critical foundation for the design and development of NaNbO3-based electronic devices.

In 1960, Megaw and co-workers systematically proposed the structural sequence of sodium niobate (NaNbO3) during cooling with consecutive phase transitions, namely, U2 (paraelectric, cubic Pm 3 ¯m) 640 ° C T2 (paraelectric, tetragonal P4/mbm) 575 ° C T1 (paraelectric, orthorhombic Ccmm) 520 ° C S (paraelectric orthorhombic Pnmm) 480 ° C R (antiferroelectric, orthorhombic Pnmm) 360 ° C P (antiferroelectric, orthorhombic Pbcm) 100 ° C N (ferroelectric, rhombohedral R3c).1–3 This widely recognized sequence has established NaNbO3 as the “most complex perovskite,” which is also considered to be an ideal model for understanding the evolution of lattice distortion in dielectrics.4,5 However, the complexity of phase transition in NaNbO3 is far more than this description. In particular, the presence of a ferroelectric Q phase with P21ma space group leads to a higher level of intricacy in both phase transition and electrical behavior under external fields. It is well known that NaNbO3 possesses an antiferroelectric P phase at room temperature, while it generally fails to exhibit the characteristic double hysteresis loop associated with antiferroelectricity. Some studies have indicated that the approximate free energies between P and Q phases give rise to the minute electric field needed by the irreversible transition from the P phase to the Q phase.6 However, based on the previous work of Zhang et al. and our research group, in situ XRD and Raman results under varying electric fields demonstrate that the required electric field is relatively higher for the transition from the P phase to the Q phase.7,8 Consequently, there is a discrepancy between macroscopic polarization measurement and microscopic structural characterization. In addition, the dielectric properties during the heating process manifest that most NaNbO3 samples display dielectric anomaly with relaxation in the temperature range of ∼100–200 °C.9 The explanation for this phenomenon is still controversial. Considering that room-temperature NaNbO3 samples usually contain an unavoidable Q phase component, some research studies propose that this anomaly arises from the incommensurate phase induced by a competition behavior during the transition from room-temperature Q phase component to P phase, while other references signify that the anomaly is just due to the transition process from the ferroelectric Q phase to the antiferroelectric P phase.10,11 Therefore, there are still many unsolved mysteries about the intricate transition between the ferroelectric and antiferroelectric phases in NaNbO3.

The phenomenological model of complex phase transition in NaNbO3 can be attributed to the synergistic effect involving the displacement of Na atoms and the distortion of NbO6 octahedra.12 Numerous works have reported on the off-center of Nb atoms and the distortion of NbO6 octahedra under external fields.4,13 Meanwhile, the influence of Na atoms cannot be overlooked, particularly during the transition between P and Q phases. According to the crystallographic structure reported by Zhang et al., NaNbO3 with the P phase exhibits two kinds of non-equivalent Na sites (denoted as Na1 and Na2), which are alternately arranged along the c axis accompanied by an antiparallel displacement between the adjacent Na1 atoms.7 This unique antiparallel displacement, in conjunction with the distortion of NbO6 octahedra, collectively forms the fourfold superlattice structure of the P phase. For comparison, both the adjacent Na1 atoms and Nb atoms present a parallel displacement in the Q phase, leading to the characteristic 1/2 superlattice reflections in electron-diffraction patterns. Of course, it is important to note that the Na atoms show constrained displacement amplitudes, combined with their inherently low atomic mass. Consequently, the traditional microstructure characterization techniques, such as x-ray diffraction, generally struggle to effectively investigate the influence of Na atoms on the phase transition of NaNbO3.12 Currently, in situ Raman spectroscopy serves as a precise probe for exploring phase transition due to its ability to sensitively detect the evolution of various phonon vibrations in response to the external physical fields.14 For example, the Raman spectrum of NaNbO3 can clearly display the splitting of phonon modes arising from two kinds of unequal Na sites, which is a hallmark feature of the P phase.15 Moreover, the ferroelectric and antiferroelectric phases, respectively, possess distinguishable characteristics in different phonon vibrations of NbO6 octahedra.16 Therefore, it is expected to reveal the influence of Na sites on the modulation principles and the related physical mechanisms of the ferroelectric-antiferroelectric transition in NaNbO3, utilizing in situ Raman spectroscopy. However, the current research in this area remains limited and requires further development.

According to the above discussion, a series of Na1+xNbO3 samples were synthesized by precisely controlling the different nonstoichiometric ratios of Na elements. The microstructure and electrical behavior could be deciphered in detail. Furthermore, the correlation between the Na atoms and the field-induced phase transition in NaNbO3 is thoroughly analyzed, based on the in situ Raman spectroscopy technique under varying electric fields and temperatures. In this work, we aim to clarify the modulation effect of Na atoms on the physical properties of NaNbO3, involving the transition from the room-temperature antiferroelectric component phase to the ferroelectric component phase and the dielectric relaxation process in the range of ∼100–200 °C. The relevant findings will provide an important foundation for further understanding the origin of electric polarization in NaNbO3, as well as the design and development of NaNbO3-based electronic devices.

Na1+xNbO3 (−0.02 ≤ x ≤ 0.02) samples were prepared by the traditional solid-state reaction method using analytical-grade Na2CO3 (99.8%) and Nb2O5 (99.9%) as starting materials. The raw materials were thoroughly dried at 150 °C and weighed according to the specified non-stoichiometric ratios. After adding anhydrous ethanol, the mixtures were ball-milled for 24 h at 200 r/min to achieve full mixing of the raw materials. Then, the resulting slurry was dried in a water bath, accompanied by transferring the powder to corundum crucibles. Calcination was conducted at 900 °C for 4 h to obtain the precursor powder. Subsequently, a secondary process under the same condition is necessary, including ball milling and drying. After adding 5 wt. % polyvinyl alcohol (PVA), the pellets were pressed at 200 MPa, with a diameter of ∼13 mm and a thickness of ∼0.5 mm. Finally, the Na1+xNbO3 ceramic series were obtained via sintering at 1300 °C for 2 h. For relevant electrical measurement, the sample thickness was reduced to ∼0.1 mm using a polishing machine, followed by high-temperature silver paste coated on both sides of the samples to form electrodes.

The crystal structure was characterized using an x-ray diffractometer with Cu Kα radiation (XRD, Rigaku Miniflex 600). Room-temperature ferroelectric hysteresis loops were measured with a Precision Premier II system (Radiant Technologies) at a frequency of 10 Hz. The dielectric properties were conducted by a polarization comprehensive measurement system (PEMS 6000, Partulab), in the temperature range of 27–627 °C. Here, the measurement frequencies are, respectively, 100 Hz, 1 kHz, 10 kHz, 50 kHz, 100 kHz, 500 kHz, and 1 MHz. In situ Raman spectra under varying electric fields and temperatures were evaluated through a high-speed and high-resolution micro-confocal Raman spectrometer using a 532 nm laser (LabRAM HR Evolution, Horiba Scientific), equipped with a temperature controller (Linkam Scientific Instruments Ltd, T96-S) and an electrometer (Keithley 6517B). The wavenumber range of Raman spectra is 50–1000 cm−1 and the groove density of gratings is 1800 gr/mm. Specifically, for the measurement of in situ Raman spectra with varying electric fields, the ceramic samples were polished and coated with silver slurry to form electrodes, after which the samples were processed into narrow strips. The measurement was performed on the section cutting plane of the strips with a direct-current bias voltage which was provided by the electrometer.

Figure 1(a) presents the XRD patterns of Na1+xNbO3 samples at room temperature, demonstrating well crystallinity with low background noise and sharp diffraction peaks. Comparison with the standard PDF card (space group Pbcm No. 01-073-0803) confirms that all samples are pure without any secondary phases, which indicates that the slight nonstoichiometric ratio of Na sites in the range of −0.02 ≤ x ≤ 0.02 does not significantly disrupt the lattice structure of parent NaNbO3. Of course, the relative intensity of the diffraction peaks varies with different Na contents. For instance, Fig. 1(b) illustrates the ratio of relative intensity between the (110) and the (100) peak (I(110)/I(100)), as a function of Na content. It is evident that the ratio initially rises rapidly before gradually approaching saturation, with the monotonic increase of x. This suggests that the variation of the Na content induces a regular distortion in the NaNbO3 lattice. To further investigate the detailed microstructure evolution, Figs. 1(c) and 1(d) provide the local enlargement of the diffraction patterns at ∼36.6° and ∼55.2°. It denotes the peaks of the fourfold superlattice in NaNbO3, as the fingerprint evidence for the presence of the antiferroelectric P phase.17 Clearly, the relative intensity of the superlattice diffraction peaks is enhanced in the Na-rich state (x > 0), indicating a preferred fourfold superlattice P phase. Conversely, the intensity of the superlattice diffraction peaks gradually weakens due to the lack of Na content, which means that the Na-deficient state (x < 0) will be unfavorable for the stability of the antiferroelectric structure in NaNbO3.

FIG. 1.

(a) Room-temperature XRD patterns of Na1+xNbO3 ceramics. (b) The variation of I(110)/I(100) with different Na contents. (c) Local enlargement of the {1 1 3/4} superlattice diffraction peak. (d) Local enlargement of the {2 1 3/4} superlattice diffraction peak.

FIG. 1.

(a) Room-temperature XRD patterns of Na1+xNbO3 ceramics. (b) The variation of I(110)/I(100) with different Na contents. (c) Local enlargement of the {1 1 3/4} superlattice diffraction peak. (d) Local enlargement of the {2 1 3/4} superlattice diffraction peak.

Close modal

In order to quantify the influence of Na-occupancy content on the lattice structure of NaNbO3 in detail, the XRD patterns of Na1+xNbO3 samples are refined using Fullprof software based on the Rietveld method. As shown in Fig. S1 (see in the supplementary material), the excellent agreement between experimental data and fitting results confirms the reliability of refinement analysis. Accordingly, the crystallographic parameters obtained from the refinement process are subsequently utilized for lattice visualization via VESTA software. Figure 2 depicts the lattice projection of all samples along the c axis. For P-phase NaNbO3, the adjacent Na1 and Na2 atoms, as well as the neighboring Nb atoms along the c axis, exhibit a relatively off-center displacement of projection on the ab plane in lattices. It is also one of the important reasons for the typical fourfold superlattice of the P phase. As can be seen from Fig. 2, the gradual decrease of Na occupancy leads to an enhanced tendency of the relatively off-center displacement of the adjacent Na atoms. Notably, the projections of the adjacent Na atoms do not overlap at all in the sample with x = −0.02. However, as the Na content decreases, the adjacent Nb atoms display an opposite trend with the weakening of the relatively off-center displacement, accompanied by a nearly complete overlapping in the sample of x = −0.02. The relevant parameters are provided in Table I. Quantitative analysis of the relatively off-center displacement of the adjacent atoms is obtained by calculating the Wyckoff position of the corresponding atoms. As a result, the displacement of the adjacent Na atoms gradually increases from 0.030 30 of x = 0.02 to 0.354 04 of x = −0.02, while that of the adjacent Nb atoms decreases from 0.029 44 of x = 0.02 to 0.005 62 of x = −0.02. Combined with the discussion of Fig. 1, the weakening of the antiferroelectric structure in Na-deficient samples can be attributed to the greater displacement of the adjacent Na atoms and the reduced displacement of the adjacent Nb atoms.

FIG. 2.

Schematic representation of Na1+xNbO3 ceramics.

FIG. 2.

Schematic representation of Na1+xNbO3 ceramics.

Close modal
TABLE I.

Refined crystallographic parameters of Na1+xNbO3 samples.

SamplesCell parametersAtom Wyckoff positionDisplacementRefinement parameters
a (Å)b (Å)c (Å)Volume (Å3)Na1Na2NbNaNbRp (%)Rwp (%)χ2
Na1.02NbO3 (x = 0.02) 5.50687 5.57043 15.53971 476.690 0.23172 0.23099 0.25533 0.03030 0.02944 5.20 8.09 2.64 
0.25000 0.21830 0.73520 
0.00000 0.25000 0.12387 
Na1.01NbO3 (x = 0.01) 5.50492 5.56872 15.54169 476.436 0.22515 0.22289 0.25386 0.03542 0.02782 5.06 7.73 2.17 
0.25000 0.20954 0.73640 
0.00000 0.25000 0.12396 
NaNbO3 (x = 0) 5.50716 5.57147 15.54145 476.853 0.27189 0.20423 0.25750 0.07482 0.02456 5.91 8.42 2.37 
0.25000 0.21807 0.73772 
0.00000 0.25000 0.12453 
Na0.99NbO3 (x = −0.01) 5.50429 5.56906 15.53971 476.351 0.12516 0.26189 0.25014 0.10084 0.0128 5.01 7.63 2.31 
0.25000 0.23976 0.74368 
0.00000 0.25000 0.12647 
Na0.98NbO3 (x = −0.02) 5.50178 5.56923 15.54559 476.328 0.33737 0.27653 0.24983 0.35404 0.00562 7.32 8.96 2.95 
0.25000 0.25396 0.75505 
0.00000 0.25000 0.12605 
SamplesCell parametersAtom Wyckoff positionDisplacementRefinement parameters
a (Å)b (Å)c (Å)Volume (Å3)Na1Na2NbNaNbRp (%)Rwp (%)χ2
Na1.02NbO3 (x = 0.02) 5.50687 5.57043 15.53971 476.690 0.23172 0.23099 0.25533 0.03030 0.02944 5.20 8.09 2.64 
0.25000 0.21830 0.73520 
0.00000 0.25000 0.12387 
Na1.01NbO3 (x = 0.01) 5.50492 5.56872 15.54169 476.436 0.22515 0.22289 0.25386 0.03542 0.02782 5.06 7.73 2.17 
0.25000 0.20954 0.73640 
0.00000 0.25000 0.12396 
NaNbO3 (x = 0) 5.50716 5.57147 15.54145 476.853 0.27189 0.20423 0.25750 0.07482 0.02456 5.91 8.42 2.37 
0.25000 0.21807 0.73772 
0.00000 0.25000 0.12453 
Na0.99NbO3 (x = −0.01) 5.50429 5.56906 15.53971 476.351 0.12516 0.26189 0.25014 0.10084 0.0128 5.01 7.63 2.31 
0.25000 0.23976 0.74368 
0.00000 0.25000 0.12647 
Na0.98NbO3 (x = −0.02) 5.50178 5.56923 15.54559 476.328 0.33737 0.27653 0.24983 0.35404 0.00562 7.32 8.96 2.95 
0.25000 0.25396 0.75505 
0.00000 0.25000 0.12605 
A more in-depth analysis of microstructural variation with varying Na content is shown by Raman spectra in Fig. 3. For room-temperature NaNbO3 with the P phase, group theoretical analysis provides the irreducible representations of the vibration modes,
(1)
where only Ag, B1g, B2g, and B3g modes are Raman active. Thus, 60 Raman active optical phonons are expected. However, the number is less in experimental Raman spectra, on account of the weak intensity and subtle frequency difference between some modes.18 Generally, the Raman modes primarily consist of phonon vibration of Na sites at lower wavenumbers and the vibration in NbO6 octahedra at higher wavenumbers.19 Particularly, the modes in the wavenumber range of ∼160–700 cm−1 are associated with the intricate phonon vibration of Nb-O bonds. Here, ν1–ν3 represent the stretching vibration of Nb–O bonds (symmetric ν1, asymmetric ν2 and ν3), while ν4–ν6 correspond to the bending vibration of Nb–O bonds (asymmetric ν4, symmetric ν5, and inactive ν6).20 Accordingly, all samples exhibit the typical P phase in the nonstoichiometric range of −0.02 ≤ x≤+0.02, characterized by the mode splitting at ∼70 cm−1. In addition, other vibration modes are also in excellent agreement with the previous reports relating to NaNbO3 with the P phase.21,22 Meanwhile, obvious difference can be observed in the Raman spectra. It is well known that the two main peaks ( ν 5 1 and ν 5 2) of the ν5 mode are highly sensitive to the electrical structure in NaNbO3. A higher relative intensity of ν 5 2 compared to ν 5 1 indicates a preferred antiferroelectric phase.23 As shown in the inset of Fig. 3, the relative intensity ratio I ( ν 5 2 ) / I ( ν 5 1 ) presents a monotonic increasing trend with increasing Na content. This proves that the Na-rich state facilitates the stabilization of the antiferroelectric P phase while the Na-deficient state would weaken the antiferroelectric structure. Moreover, an evident variation is also reflected by the relative intensity ratio I(ν2)/I(ν1) of the asymmetric stretching ν2 and the symmetric stretching ν1 in Nb–O bonds. The ratio of I(ν2)/I(ν1) decreases as the Na content deviates from the stoichiometric ratio, suggesting that a distortion appears in NbO6 octahedra with the variation of the Na content. Thus, the room-temperature XRD and Raman results demonstrate that the nonstoichiometric design of Na sites can induce quantifiable crystal distortion while preserving the antiferroelectric P phase framework of NaNbO3. Specifically, the Na-rich Na1+xNbO3 samples tend to stabilize a more perfect P phase, whereas the Na-deficient samples will develop an octahedral distortion within the preserved P-phase matrix, although no additional phase is formed with nonstoichiometric regulation.
FIG. 3.

Raman spectra of Na1+xNbO3 ceramics at room temperature. The inset is the variation of the relative intensity ratio for I ( ν 5 2 ) / I ( ν 5 1 ) and I(ν2)/I(ν1).

FIG. 3.

Raman spectra of Na1+xNbO3 ceramics at room temperature. The inset is the variation of the relative intensity ratio for I ( ν 5 2 ) / I ( ν 5 1 ) and I(ν2)/I(ν1).

Close modal

Distinct from the conventional antiferroelectrics, NaNbO3 is characterized by its irreversible transition from the antiferroelectric P phase to the ferroelectric Q phase under electric fields, which typically manifests as square-shaped ferroelectric-like hysteresis loops.24 To investigate the modulating effect of Na-site occupancy on field-induced phase transition, the room-temperature polarization-electric field hysteresis loops of Na1+xNbO3 are presented in Fig. 4. Previous studies have revealed that the synthesis of polycrystalline NaNbO3 would inevitably introduce substantial vacancy defects, mainly involving Na vacancies and oxygen vacancies with traces of Nb vacancies.25,26 Notably, the pronounced volatility of Na at high temperatures leads to that the vacancy effect cannot be negligible in NaNbO3.27 This inherent limitation has historically constrained the maximum of the applicable electric field during the polarization measurement of NaNbO3.28 Consistent with these observations, the parent NaNbO3 in this work exhibits a breakdown field of ∼90 kV/cm. Intriguingly, controlled deviations from Na-site stoichiometry, including both Na-rich and Na-deficient compositions, result in a significant enhancement of the breakdown electric field reaching ∼140 kV/cm. This counterintuitive phenomenon, particularly the improved breakdown field observed in Na-deficient Na1+xNbO3 samples, suggests that the low breakdown field in general NaNbO3 may not originate solely from vacancy effects. It could also potentially correlate with the intrinsic structural characteristics, such as the crystal distortion in NbO6 octahedra. Of course, the underlying mechanism requires further elucidation. As shown in Fig. 4, the shape of hysteresis loops still demonstrates a regular trend despite the disparity in the maximum applicable field between the parent NaNbO3 and the modified samples. The Na-rich samples tend to display an enhanced squareness of hysteresis loops at room temperature, indicative of the preferred ferroelectric feature. Conversely, a progressive weakening of the ferroelectric response is observed with decreasing Na content, as evidenced by the gradual reduction in loop rectangularity.

FIG. 4.

The room-temperature polarization–electric field hysteresis loops of Na1+xNbO3 ceramics.

FIG. 4.

The room-temperature polarization–electric field hysteresis loops of Na1+xNbO3 ceramics.

Close modal

Microstructural analysis reveals that the Na-rich samples possess enhanced stabilization of the antiferroelectric structure while the Na-deficient state leads to a progressive degradation of the antiferroelectric structure. However, hysteresis loop measurement indicates that those samples with reinforced antiferroelectric structures (x > 0) are easier for triggering the transition to the ferroelectric state, while those with weakened antiferroelectric stability show suppressed trigger response to the electric field. This behavior is interesting but seems to be counterintuitive. To elucidate this anomaly, as well as the relating origin, the room-temperature in situ Raman spectra are measured under varying electric fields, as shown in Fig. 5. For clarity, the two samples with x = −0.02 and 0.02 are, respectively, chosen due to the most obvious differences in their structure and performance. Both samples present distinct three-stage structural evolution of the phonon modes under the applied electric field. During the initial stage with lower electric field (E < E1), all vibration modes are insensitive to the field without any detectable response, corresponding to the stable antiferroelectric state (AFE). Upon reaching the first critical field (E1), abrupt variations in different phonon modes mainly manifest through three concurrent aspects, including the merging of phonon modes for two kinds of different Na sites, the dramatic intensity inversion between ν 5 2 (sharp decrease) and ν 5 1 (rapid enhancement), as well as the complete disappearance of the ν2 mode accompanied by a significant intensity reduction and a wavenumber shift of the ν1 mode. These collective evolution indicate the partial breakdown of the fourfold superlattice configuration of P-phase NaNbO3 in the second stage after the electric field exceeds E1 (E1 < E < E2), marking the trend of transition to the Q phase. Notably, residual signatures of the P phase persist during this stage, suggesting a superposition of the antiferroelectric and ferroelectric states (AFE + FE) in NaNbO3. When exceeding the second critical field (E2), the samples undergo further crystal evolution, involving the complete merging of two kinds of Na-site phonon modes, the gradual stabilization of ν 5 2 intensity, and the secondary wavenumber shift in ν5 and ν1 modes. Therefore, a full transition to the ferroelectric (FE) Q state is completed of NaNbO3 in the third stage (E > E2), evidenced by field-saturated phonon responses beyond E2.

FIG. 5.

The in situ Raman spectra under varying electric fields of Na1+xNbO3 with x = −0.02 (a) and (c) and 0.02 (b) and (d). Here, FE and AFE are, respectively, short for “ferroelectric” and “antiferroelectric.” For the sample with x = −0.02, the values of E1 and E2 are, respectively, 55 and 60 kV/cm, while the values of E1 and E2 are, respectively, 36 and 48 kV/cm for the sample with x = 0.02.

FIG. 5.

The in situ Raman spectra under varying electric fields of Na1+xNbO3 with x = −0.02 (a) and (c) and 0.02 (b) and (d). Here, FE and AFE are, respectively, short for “ferroelectric” and “antiferroelectric.” For the sample with x = −0.02, the values of E1 and E2 are, respectively, 55 and 60 kV/cm, while the values of E1 and E2 are, respectively, 36 and 48 kV/cm for the sample with x = 0.02.

Close modal

Moreover, a comparative analysis of in situ Raman spectra with varying electric fields between the samples with x = −0.02 and 0.02 reveals a markedly improved critical fields in the Na-deficient state, which is consistent with the polarization hysteresis characteristics in Fig. 4. This correlation reaffirms that NaNbO3 with an optimized antiferroelectric P phase structure will be more easily triggered to the Q phase driven by an electric field. These findings conclusively demonstrate the critical modulatory role of Na-site occupancy in governing the field-induced phase transition behavior of NaNbO3 systems. Crystallographic analysis indicates that the antiferroelectric structure collectively arises from the antiparallel displacement of the adjacent Na1 atoms and the distorted NbO6 octahedra, in the prototypical P phase.29 However, both polarization hysteresis and in situ spectra suggest that this “ideal” P phase exhibits intrinsic metastability under electric fields, readily transitioning to the ferroelectric Q state. Na deficiency results in the increase of the relatively off-center displacement of the adjacent Na atoms, coupled with the concomitant crystal distortion of NbO6 octahedra, to a certain extent. The electric field-induced transition from the P phase to the Q phase involves transforming the antiparallel displacement of the adjacent Na1 atoms into parallel displacement. Consequently, the increase of the relatively off-center displacement of the adjacent Na atoms necessarily demands higher energy provided by the electric field to realize this phase transition process, thereby increasing the difficulty of the field-induced phase transition. Thus, it is a remarkable phenomenon that enhanced antiferroelectric structure facilitates to stronger ferroelectric characteristics at room temperature, from the macroscopic perspective.

After exploring the modulation influence of Na-site occupancy on the field-induced phase transition in NaNbO3, the effect on temperature-dependent phase evolution will be further investigated. Figures 6(a)6(e) present the temperature dependence of the dielectric constant (ε) of Na1+xNbO3 ceramics in the temperature range of 27–627 °C. All samples display a frequency-independent peak near 400 °C, corresponding to the transition from the P phase to the R phase in NaNbO3. It is well consistent with the previous report.30  Figure 6(f) shows the temperature-dependent dielectric loss for all samples, accompanied by a pronounced abrupt variation at ∼400 °C, which provides additional compelling evidence for the first-order phase transition. Furthermore, dielectric characterization reveals another anomaly (∼100 °C) with a diffuse behavior in the dielectric constant, which indicates dielectric relaxation. For clarity, the corresponding local enlargements are plotted in the insets of Figs. 6(a)6(e). The relaxation is gradually prominent with the increasing Na content. It typically arises from the presence of the disordered state rather than first-order phase transitions.31,32

FIG. 6.

(a)–(e) Temperature dependence of the dielectric constant for Na1+xNbO3 ceramics. (f) Temperature dependence of dielectric loss for Na1+xNbO3 ceramics at the fixed frequency of 10 kHz.

FIG. 6.

(a)–(e) Temperature dependence of the dielectric constant for Na1+xNbO3 ceramics. (f) Temperature dependence of dielectric loss for Na1+xNbO3 ceramics at the fixed frequency of 10 kHz.

Close modal

The different relaxations at ∼100 °C for Na1+xNbO3 samples may potentially correlate with the different structural evolution tendency during heat processing, due to the slight difference in room-temperature microstructure. Thus, it is a primary issue to clarify. According to the curves in Fig. 6, the relaxation process for all samples has completely finished upon reaching ∼200 °C. Consequently, the Raman spectra at 200 °C are conducted and presented in Fig. 7(a). It is evident that all spectra exhibit negligible difference in both mode shape and relative intensity, suggesting that the phase transition pathway during the heating process is fundamentally consistent for the Na1+xNbO3 samples, despite the subtle structural discrepancy at room temperature. In addition, it should be noted that the difference is significant in phonon vibration, respectively, at room temperature and 200 °C, though extensive studies have demonstrated the P phase of NaNbO3 at 200 °C during the whole phase transition sequence.33 That is, the P-phase lattice at room temperature will inevitably undergo a gradual lattice expansion and bond distortion with increasing temperature, in spite of the same space group and crystal symmetry. Hence, for Na1+xNbO3 series, the lattice structure at 200 °C can be regarded as a distorted P phase. For clarity, in subsequent discussion, the structure state at 200 °C is designated as the Ph state (P phase at high temperatures). The phonon vibration of Na1+xNbO3 samples evolves from a distinguishable mode characteristic of room-temperature P phase to a nearly identical feature of the Ph state at 200 °C. The discrepancy of structure evolution under varying temperatures is likely the essential reason for the relaxation behavior during the process from the P state to the Ph state.

FIG. 7.

(a) Raman spectra of Na1+xNbO3 ceramics at 200 °C. (b) and (c) In situ Raman spectra of Na1+xNbO3 with x = −0.02 (b) and 0.02 (c) accompanied by the temperature increasing from room temperature to 200 °C.

FIG. 7.

(a) Raman spectra of Na1+xNbO3 ceramics at 200 °C. (b) and (c) In situ Raman spectra of Na1+xNbO3 with x = −0.02 (b) and 0.02 (c) accompanied by the temperature increasing from room temperature to 200 °C.

Close modal

In order to further analyze this evolution discrepancy, the two samples with x = −0.02 and 0.02 are selected to measure the temperature-dependent Raman spectra from room temperature to 200 °C, due to their most different behaviors of relaxation. As shown in Figs. 7(b) and 7(c), all phonon modes exhibit a slight redshift with increasing temperature due to the lattice expansion under the thermal effect.34 Moreover, the distinct response of various phonon vibrations to varying temperatures can be directly observed, involving the gradual disappearance of splitting behavior for Na-site phonon modes, the progressive merging of NbO6 rotation (denoted as NbO6 rot in Fig. 3) mode, and the intensity redistribution of different ν5 modes. The increasing temperature provides additional energy to lattices, leading to structural distortion compared to that at room temperature. The comparison between Figs. 7(b) and 6(c) reveals that the corresponding characteristic temperatures associated with different mode evolution are clearly lower in the sample with x = −0.02 than those in the sample with x = 0.02, during the heating process. This indicates the lower energy needed by the sample with x = −0.02 for the structural evolution from the P state to the Ph state. Relevant study has demonstrated that the gradual decrease of temperature will enhance the relatively off-center displacement of the Na atoms as well as the Nb atoms in NaNbO3, which reflects that the lattice structure can present more pronounced symmetry breaking at lower temperatures.13 Thus, compared with the room-temperature P phase, the Ph state at 200 °C should exhibit weaker displacement including both Na and Nb atoms. The analysis related to the above figures indicates that the Na deficiency in NaNbO3, indeed, induces the distortion of the P phase. Namely, Na-deficient Na1+xNbO3 samples possess stronger displacement of the adjacent Na atoms but weaker displacement of the adjacent Nb atoms, compared to the perfect P phase. Hence, it is clear that the Na-deficient P phase at room temperature and the Ph state at 200 °C display consistent behavior of weak displacement for Nb atoms, accompanied by a significant difference in the displacement of Na atoms. Accordingly, for the sample with x = −0.02, a lower energy requirement during the evolution from the room-temperature P phase to the Ph state signifies that, the Na-deficient P phase at room temperature and the Ph state originate from the same mechanism during the phase evolution process from the perfect room-temperature P phase. That is, the variation of displacement behavior in the Nb atoms plays a dominating role during the phase transition from room temperature to 200 °C. Correspondingly, the Na-rich samples demonstrate a more stable P phase at room temperature accompanied by a stronger displacement of the Nb atoms, thereby requiring higher energy to evolve into the Ph state.

In addition, compared with the sample with x = 0.02, the evolution behavior of various phonons in the sample with x = −0.02 is relatively smoother during the whole temperature range. This is another manifestation of the lower energy required in the evolution of the sample with x = −0.02. Thus, it can be concluded that the relatively off-center displacement of the adjacent Nb atoms is the critical factor during the phase evolution in NaNbO3 from room temperature to 200 °C. The distorted NaNbO3 at ∼200 °C exhibits a more similar displacement of the Nb atoms to that of room-temperature NaNbO3 with Na-deficient state, which belongs to a metastable antiferroelectric structure. Of course, although the direct influence of Na atoms on this phase evolution is relatively minor, the variation of Na content can indirectly have an impact on the displacement behavior of Nb atoms, thereby significantly modulating the relaxation behavior at ∼100 °C. It can be attributed to the evolution of the typical antiferroelectric P phase at room temperature to the metastable distorted P phase at higher temperatures.

In summary, a series of nonstoichiometric Na1+xNbO3 ceramics are successfully synthesized to investigate the regulation mechanism of Na-site ions on the field-induced phase transition in NaNbO3. On the basis of the superior capability of the Raman spectroscopy technique in detecting phase transition details, the microstructure characterization of Na1+xNbO3 is systematically analyzed via in situ Raman spectra under varying temperatures and electric fields. The modulation of Na-site occupancy on field-induced phase transition is reflected by easier triggering to the Q phase of NaNbO3 with an ideal antiferroelectric P phase. Combined with the related ferroelectric and dielectric properties, in situ Raman results confirm that the Na-rich Na1+xNbO3 samples tend to stabilize a more perfect P phase at room temperature, whereas the Na-deficient samples will develop an octahedral distortion in the preserved P-phase matrix. The distorted NaNbO3 at ∼200 °C exhibits a more similar displacement of Nb atoms to that of room-temperature NaNbO3 with a Na-deficient state, which belongs to a metastable antiferroelectric structure. Moreover, the regulation of Na-site atoms on the relaxation behavior at ∼100 °C should be attributed to the evolution of the typical antiferroelectric P phase at room temperature to the the distorted P phase at higher temperatures, through indirectly having impact on the displacement behavior of the Nb atoms. The findings in this work provide an additional design methodology for the structure–activity relationship of NaNbO3-based materials.

See the supplementary material for the refinement of Na1+xNbO3 ceramics.

This work was supported by the Natural Science Foundation of Guangxi Province (No. 2024GXNSFAA010295), Middle-aged and Young Teachers' Basic Ability Promotion Project of Guangxi (Nos. 2024KY1887 and 2023KY0065), and Innovation Project of Guangxi Graduate Education (No. YCSW2024156).

The authors have no conflict of interest.

H. Cui and X.X.H. have contributed equally to this work.

H. Cui: Investigation (equal); Writing – original draft (equal). X. X. Huang: Funding acquisition (equal); Investigation (equal); Writing – original draft (equal). X. L. Jiang: Investigation (equal); Project administration (equal). C. M. Zhu: Conceptualization (equal); Writing – review & editing (equal). L. G. Wang: Funding acquisition (equal); Writing – review & editing (equal). R. Wang: Validation (equal). S. Lu: Validation (equal). G. B. Yu: Funding acquisition (equal); Software (lead).

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

1.
C.
Darlington
and
H.
Megaw
, “
The low-temperature phase transition of sodium niobate and the structure of the low temperature phase N
,”
Acta Crystallogr. Sect. B Struct. Crystallogr. Crystal Chem.
29
,
2171
(
1973
).
2.
Y.
Tian
,
L.
Li
,
J.
Geng
,
H. B.
Shi
,
Y. H.
Xu
,
W. J.
Shi
,
X.
Guo
,
L. N.
She
,
L.
Jin
,
W. Y.
Ge
, and
X. Y.
Wei
, “
Structure evolution and electric-field-induced reversible transition in perovskite Na(Nb1−xTax)O3 ceramics
,”
J. Am. Ceram. Soc.
106
,
5907
5921
(
2023
).
3.
S.
Bin Anooz
,
Y.
Wang
,
P.
Petrik
,
M.
de Oliveira Guimaraes
,
M.
Schmidbauer
, and
J.
Schwarzkopf
, “
High temperature phase transitions in NaNbO3 epitaxial films grown under tensile lattice strain
,”
Appl. Phys. Lett.
120
,
202901
(
2022
).
4.
M.
Jauhari
,
S. K.
Mishra
,
R.
Mittal
,
P. U.
Sastry
, and
S. L.
Chaplot
, “
Effect of chemical pressure on competition and cooperation between polar and antiferrodistortive distortions in sodium niobate
,”
Phys. Rev. Mater.
1
,
074411
(
2017
).
5.
C. S.
Htet
,
S.
Nayak
,
A.
Manjón-Sanz
,
J.
Liu
,
J.
Kong
,
D. R.
Sørensen
,
F.
Marlton
,
M. R. V.
Jørgensen
, and
A.
Pramanick
, “
Atomic structural mechanism for ferroelectric-antiferroelectric transformation in perovskite NaNbO3
,”
Phys. Rev. B
105
,
174113
(
2022
).
6.
G.
Luo
,
K.
Yang
,
X.
Fu
,
L.
Ma
,
Z.
Che
,
X.
Chen
,
Z.
Cen
,
Q.
Feng
, and
N.
Luo
, “
Unveiling the role of chemical pressure on antiferroelectricity in NaNbO3-based antiferroelectric ceramics
,”
J. Appl. Phys.
134
,
234101
(
2023
).
7.
M. H.
Zhang
,
C.
Zhao
,
L.
Fulanović
,
J.
Rödel
,
N.
Novak
,
A.
Schökel
, and
J.
Koruza
, “
Revealing the mechanism of electric-field-induced phase transition in antiferroelectric NaNbO3 by in situ high-energy x-ray diffraction
,”
Appl. Phys. Lett.
118
,
132903
(
2021
).
8.
L. G.
Wang
,
X. L.
Jiang
,
C. M.
Zhu
,
G. B.
Yu
,
X. F.
Su
,
M. Y.
Qin
,
S.
Lu
,
N.
Shen
, and
X. X.
Zheng
, “
Visualization of the phonon evolution behavior in NaNbO3 single crystal during field-induced phase transition by in situ Raman spectroscopy
,”
Appl. Phys. Lett.
126
,
012905
(
2025
).
9.
J.
Koruza
,
H.
Tellier
,
B.
Malič
,
V.
Bobnar
, and
M.
Kosec
, “
Phase transitions of sodium niobate powder and ceramics, prepared by solid state synthesis
,”
J. Appl. Phys.
108
,
113509
(
2010
).
10.
H.
Guo
,
H.
Shimizu
, and
C. A.
Randall
, “
Direct evidence of an incommensurate phase in NaNbO3 and its implication in NaNbO3-based lead-free antiferroelectrics
,”
Appl. Phys. Lett.
107
,
112904
(
2015
).
11.
L. A.
Reznichenko
,
L. A.
Shilkina
,
E. S.
Gagarina
,
I. P.
Raevskii
,
E. A.
Dul’kin
,
E. M.
Kuznetsova
, and
V. V.
Akhnazarova
, “
Structural instabilities, incommensurate modulations and P and Q phases in sodium niobate in the temperature range 300–500 K
,”
Crystallogr. Rep.
48
,
448
456
(
2003
).
12.
J.
Koruza
,
P.
Groszewicz
,
H.
Breitzke
,
G.
Buntkowsky
,
T.
Rojac
, and
B.
Malič
, “
Grain-size-induced ferroelectricity in NaNbO3
,”
Acta Mater.
126
,
77
85
(
2017
).
13.
L.
Jiang
,
D. C.
Mitchell
,
W.
Dmowski
, and
T.
Egami
, “
Local structure of NaNbO3: A neutron scattering study
,”
Phys. Rev. B
88
,
014105
(
2013
).
14.
Y.
Ye
,
A.
Cui
,
L.
Gao
,
K.
Jiang
,
L.
Zhu
,
J.
Zhang
,
L.
Shang
,
Y.
Li
,
G.
Wang
,
X.
Dong
,
Z.
Hu
, and
J.
Chu
, “
In situ Raman scattering studies of pressure-temperature phase diagrams in antiferroelectric xCaSnO3-modified NaNbO3 ceramics
,”
Appl. Phys. Lett.
119
,
132905
(
2021
).
15.
M.
Jauhari
,
S. K.
Mishra
,
J. K.
Poswal
,
R.
Mittal
, and
S. L.
Chaplot
, “
Evidence of low-temperature phase transition in BaTiO3-modified NaNbO3: Raman spectroscopy study
,”
J. Raman Spectrosc.
50
,
1949
1955
(
2019
).
16.
R. A.
Shakhovoy
,
S. I.
Raevskaya
,
L. A.
Shakhovaya
,
D. V.
Suzdalev
,
I. P.
Raevski
,
Y. I.
Yuzyuk
,
A. F.
Semenchev
, and
M.
El Marssi
, “
Ferroelectric Q and antiferroelectric P phases’ coexistence and local phase transitions in oxygen-deficient NaNbO3 single crystal: Micro-Raman, dielectric and dilatometric studies
,”
J. Raman Spectrosc.
43
,
1141
1145
(
2012
).
17.
L.
Gao
,
H.
Guo
,
S.
Zhang
, and
V. A.
Randall
, “
Stabilized antiferroelectricity in xBiScO3-(1−x)NaNbO3 lead-free ceramics with established double hysteresis loops
,”
Appl. Phys. Lett.
112
,
092905
(
2018
).
18.
K. K.
Mishra
,
V.
Sivasubramanian
, and
A. K.
Arora
, “
Low-temperature Raman spectroscopic studies in NaNbO3
,”
J. Raman Spectrosc.
42
,
517
521
(
2011
).
19.
G. F.
Teixeira
,
H. S.
Seleghini
,
W. B.
Bastos
,
N.
Jacomaci
,
B.
Stojadinović
,
Z.
Dohčević-Mitrović
,
F.
Colmati
,
N. A.
San-Miguel
,
E.
Longo
, and
M. A.
Zaghete
, “
On the coexistence of ferroelectric and antiferroelectric polymorphs in NaNbO3 fibers at room temperature
,”
J. Mater. Chem. C
11
,
5524
5533
(
2023
).
20.
N. H.
Khansur
,
U. R.
Eckstein
,
M.
Bergler
,
A.
Martin
,
K.
Wang
,
J. F.
Li
,
M. R.
Cicconi
,
K.
Hatano
,
K.
Kakimoto
,
D.
de Ligny
, and
K. G.
Webber
, “
In situ combined stress- and temperature-dependent Raman spectroscopy of Li-doped (Na, K)NbO3
,”
J. Am. Ceram. Soc.
105
,
2735
2743
(
2022
).
21.
H.
Qi
,
G.
Wang
,
Y. C.
Zhang
,
D. W.
Wang
,
H.
Liu
,
S. Q.
Deng
,
R. Z.
Zuo
, and
J.
Chen
, “
Tunable phase structure in NaNbO3 ceramics by grain-size effect, electric field and heat treatment
,”
Acta Mater.
248
,
118778
(
2023
).
22.
M.
Jauhari
,
S. K.
Mishra
,
R.
Mittal
, and
S. L.
Chaplot
, “
Probing of structural phase transitions in barium titanate modified sodium niobate using Raman scattering
,”
J. Raman Spectrosc.
50
,
1177
1185
(
2019
).
23.
J.
Ye
,
G.
Wang
,
X.
Chen
,
F.
Cao
, and
X.
Dong
, “
Enhanced antiferroelectricity and double hysteresis loop observed in lead-free (1−x)NaNbO3-xCaSnO3 ceramics
,”
Appl. Phys. Lett.
114
,
122901
(
2019
).
24.
F.
Li
,
K.
Li
,
M.
Long
,
C.
Wang
,
G.
Chen
, and
J.
Zhai
, “
Ferroelectric-relaxor crossover induce large electrocaloric effect with ultrawide temperature span in NaNbO3-based lead-free ceramics
,”
Appl. Phys. Lett.
118
,
043902
(
2021
).
25.
L.
Villa
and
K.
Albe
, “
Role of doping and defect quenching in antiferroelectric NaNbO3 from first principles
,”
Phys. Rev. B
106
,
134101
(
2022
).
26.
M. H.
Zhang
,
H.
Ding
,
S.
Egert
,
C.
Zhao
,
L.
Villa
,
L.
Fulanović
,
P. B.
Groszewicz
,
G.
Buntkowsky
,
H.
Kleebe
,
K.
Albe
,
A.
Klein
, and
J.
Koruza
, “
Tailoring high-energy storage NaNbO3-based materials from antiferroelectric to relaxor states
,”
Nat. Commun.
14
,
1525
(
2023
).
27.
L.
Chao
,
Y.
Hou
,
M.
Zheng
, and
M.
Zhu
, “
High dense structure boosts stability of antiferroelectric phase of NaNbO3 polycrystalline ceramics
,”
Appl. Phys. Lett.
108
,
212902
(
2016
).
28.
L.
Chao
,
Y.
Hou
,
M.
Zheng
,
Y.
Yue
, and
M.
Zhu
, “
Macroscopic ferroelectricity and piezoelectricity in nanostructured NaNbO3 ceramics
,”
Appl. Phys. Lett.
110
,
122901
(
2017
).
29.
S. K.
Mishra
,
R.
Mittal
,
V. Y.
Pomjakushin
, and
S. L.
Chaplot
, “
Phase stability and structural temperature dependence in sodium niobate: A high-resolution powder neutron diffraction study
,”
Phys. Rev. B
83
,
134105
(
2011
).
30.
L.
Wang
,
X.
Huang
,
C.
Zhu
,
X.
Su
,
G.
Yu
,
P.
Zeng
,
C.
Jiang
, and
Y.
Wang
, “
Modulation on structure and electric properties by tantalum of lead-free NaNbO3 ceramics
,”
J. Raman Spectrosc.
55
,
1010
1018
(
2024
).
31.
I.
Levin
,
F.
Yang
,
R.
Maier
,
W. J.
Laws
,
D. S.
Keeble
,
G.
Cibin
, and
D. C.
Sinclair
, “
Displacive order-disorder behavior and intrinsic clustering of lattice distortions in Bi-substituted NaNbO3
,”
Adv. Funct. Mater.
30
,
2001840
(
2020
).
32.
S. K.
Mishra
,
M.
Jauhari
,
R.
Mittal
,
P. S. R.
Krishna
,
V. R.
Reddy
, and
S. L.
Chaplot
, “
Evidence for existence of functional monoclinic phase in sodium niobate based solid solution by powder neutron diffraction
,”
Appl. Phys. Lett.
112
,
182905
(
2018
).
33.
Y.
Yang
,
B.
Xu
,
C.
Xu
,
W.
Ren
, and
L.
Bellaiche
, “
Understanding and revisiting the most complex perovskite system via atomistic simulations
,”
Phys. Rev. B
97
,
174106
(
2018
).
34.
L.
Wang
,
H.
Qin
,
C.
Zhu
,
C.
Jiang
,
G.
Yu
,
X.
Su
, and
X.
Jiang
, “
Identifying the opposite shift behavior between depolarization and relaxation in BNT-based ceramics
,”
J. Phys. Chem. C
128
,
13199
13206
(
2024
).