Fact-checked by Grok 2 weeks ago

Electrostriction

Electrostriction is a universal electromechanical phenomenon observed in all dielectric materials, characterized by a quadratic strain response proportional to the square of the applied electric field strength, resulting in deformation that is independent of the field's direction or polarity. This effect arises from the Coulombic attraction between polarized charges within the material's lattice, leading to a compression or expansion of the dielectric structure under the electric field, with the induced strain often expressed through fourth-rank tensor coefficients such as the electrostriction coefficient Q or M. In contrast to piezoelectricity, which produces a linear strain directly proportional to the electric field and requires non-centrosymmetric crystal structures, electrostriction is inherently nonlinear, exhibits no hysteresis in ideal cases, and occurs universally across insulators, including centrosymmetric crystals, polymers, and ferroelectrics. Electrostriction coefficients vary widely by material class: for example, typical values of Q range from $10^{-3} m⁴/C² in ferroelectrics like (PZT) to as high as $10^{3} m⁴/C² in polymers, enabling strains that oscillate at twice the frequency of an alternating current field. Applications leverage this property for high-precision devices, including (MEMS) actuators, ultrasonic transducers, and micropumps, where the anhysteretic response ensures reproducibility and suitability for low-dimensional structures. Recent advancements have highlighted "giant" electrostriction in novel materials, such as gadolinium-doped ceria and lead halide perovskites, achieving effective piezoelectric coefficients up to d_{33} \sim 2 \times 10^5 pC/N—orders of magnitude larger than conventional piezoelectrics—sparking renewed interest since its notable rediscovery in polymers like PVDF-TrFE in 1998. These developments focus on lead-free alternatives, such as doped bismuth oxides and relaxor ferroelectrics, to address environmental concerns and enhance performance in fields like medical devices, sensors, and . As of 2025, further progress includes enhanced electrostriction in Sm-doped ceria, W-Mg co-doped sodium bismuth titanate, nanodomain-engineered relaxor ferroelectric polymers, and substituted molybdates like , alongside textured .

Definition and Basics

Definition

Electrostriction is the mechanical deformation, or , exhibited by all materials when subjected to an applied , where the is proportional to the square of the field strength. This effect arises from electrostatic interactions within the material and is a of insulators, occurring regardless of the material's crystal . Unlike effects that depend on specific structural asymmetries, electrostriction manifests in every , making it a fundamental electromechanical coupling inherent to these materials. The strain induced by electrostriction is an even function of the , meaning the deformation direction and magnitude remain the same whether the field is positive or negative, due to the quadratic dependence on the field. This contrasts with the linear piezoelectric effect, which produces strain directly proportional to the field and is limited to non-centrosymmetric crystals. The magnitude of electrostrictive strain in a given material is quantified by the electrostrictive coefficient Q, a material-specific tensor that relates the strain components to the square of the electric polarization. This coefficient provides a measure of the effect's strength, varying across different dielectrics based on their atomic and molecular structure.

Historical Background

In 1881, French physicist Gabriel Lippmann mathematically predicted electrostriction as a quadratic electromechanical effect inherent to all dielectrics, positioning it as the nonlinear counterpart to the linear piezoelectric effect he had deduced from thermodynamic principles earlier that year. Concurrently, German physicist derived a theoretical framework for electrostriction in compressible fluids, linking it to thermodynamic principles and considerations. The Curie brothers, and , reported initial observations of electrostrictive deformations in both liquids and crystalline solids during their 1880s investigations into , noting quadratic strain responses in materials like under . German physicist Friedrich Pockels advanced these findings in 1885 through systematic studies of elasto-optic and electro-optic responses, establishing the presence of such effects across diverse solid structures. Theoretical development accelerated in the and , with Warren P. Mason formulating key models linking electrostriction to dielectric permittivity and material compliance, particularly in ferroelectric ceramics like (BaTiO₃). Mason's 1949 analysis demonstrated how electrostrictive scales with the square of , providing a framework for predicting behavior in polycrystalline materials and integrating it with Maxwell's stress tensor for electromagnetic forces. This period solidified electrostriction's role in electromechanical , influencing wartime applications in ultrasonics. By the 1950s, experimental confirmations distinguished electrostriction from converse through precise measurements in relaxor ferroelectrics and ceramics, such as (PZT), where nonlinear strain and field-squared dependence were isolated using biasing techniques. Researchers like Bertram Jaffe verified these effects in high-permittivity materials, confirming electrostriction's universality and non-polar nature, even in centrosymmetric dielectrics lacking , through interferometric and methods. These milestones established electrostriction as a fundamental, quadratic electromechanical coupling.

Theoretical Foundations

Physical Mechanism

Electrostriction arises from the interaction of an applied with the material, inducing a that generates temporary dipoles even in non-polar substances. These induced dipoles experience attractive and repulsive forces, leading to a redistribution of charges within the material and resulting in anisotropic or . In non-polar , the polarizes atoms or molecules, creating aligned dipoles whose mutual interactions minimize the system's electrostatic , thereby deforming the or molecular structure. A key contributor to this deformation is the Maxwell stress, which manifests as electrostatic forces acting on the material's surface due to and on the bulk from gradients in the . On the surface, attracts opposite charges across the interface, compressing the material perpendicular to direction. In the bulk, spatial variations in produce additional body forces that further distort the structure, enhancing the overall . These stresses are present in all dielectrics and operate independently of any permanent . The mechanism varies by material class. In polymers, such as those based on , the aligns polarizable chains, reorienting them along the field and causing macroscopic through entropic and conformational changes. In ceramics, like lead magnesium niobate, the effect stems from lattice distortions where the field displaces ions, altering interatomic distances and bond angles in the . This effect is inherently because the is linearly proportional to the strength, while the resulting depends on the product of polarization components, scaling with the square of the field (E²). This quadratic dependence reflects the second-order energy minimization in the presence of the field, distinguishing electrostriction from linear effects like , which rely on pre-existing permanent dipoles.

Mathematical Description

The mathematical description of electrostriction is grounded in the constitutive relations that link mechanical to electric or in materials. The fundamental expresses the tensor S_{ij} as a of the \mathbf{P}: S_{ij} = Q_{ijkl} P_k P_l where Q_{ijkl} is the fourth-rank electrostrictive tensor, also known as the polarization-squared electrostriction tensor, with units of m⁴/C². This relation captures the universal electrostrictive response observed in all , independent of , due to the even-order dependence on . The tensor Q_{ijkl} possesses intrinsic symmetries arising from the symmetry of the strain tensor (S_{ij} = S_{ji}) and the quadratic polarization term (P_k P_l = P_l P_k), reducing the number of potentially independent components from 81 to 36 in the most general case without additional material symmetries. In , this is often represented as a 6×6 relating the six components to the squares and products of the three components. For materials with higher , such as cubic , the tensor simplifies to three independent components: Q_{11}, Q_{12}, and Q_{44}. In isotropic materials, these further reduce to two independent parameters, with the shear component satisfying Q_{44} = \frac{Q_{11} - Q_{12}}{2} to preserve rotational invariance. An equivalent formulation expresses the strain in terms of the \mathbf{E}: S_{ij} = M_{ijkl} E_k E_l where M_{ijkl} is the electrostrictive tensor relating to the field squared, with units of m²/V². For linear dielectrics, where is linearly related to the field via P_k = \epsilon_0 \chi_{kl} E_l (with \epsilon_0 the and \chi_{kl} the tensor), the two tensors are connected by M_{ijkl} = Q_{ijmn} (\epsilon_0 \chi_{mk}) (\epsilon_0 \chi_{nl}). This yields a field dependence of S \propto \epsilon E^2, where \epsilon = \epsilon_0 (1 + \chi) is the tensor, highlighting the quadratic scaling with field strength and the role of properties in amplifying the effect. These relations emerge from the thermodynamic framework governing electromechanical coupling. The density G = U - TS - \mathbf{E} \cdot \mathbf{P} (where U is , T , and S ) is expanded in powers of S_{ij} and E_k. The electrostrictive contribution appears as a third-order term in the expansion, \Delta G = -\frac{1}{2} m_{ijkl} S_{ij} E_k E_l, where m_{ijkl} relates to the field squared. The is then obtained as S_{ij} = s_{ijkl} \sigma_{kl} + M_{ijkl} E_k E_l, with s_{ijkl} the tensor, linking electrostriction to the second of the with respect to field and . This thermodynamic derivation also connects the electrostrictive coefficients to of with respect to , Q_{ijkl} \propto \frac{\partial \epsilon_{kl}}{\partial S_{ij}}, providing a basis for computational prediction from response.

Versus Piezoelectricity

Electrostriction occurs universally in all materials, irrespective of their , as it arises from the interaction between the and the material's response. In contrast, is confined to non- materials, where the lack of inversion allows for a between and mechanical without requiring a quadratic dependence. The electrostrictive effect produces a strain that is quadratic with respect to the applied electric field, resulting in an even response function that always yields positive strain regardless of field polarity and inherently lacks hysteresis, promoting reversible and stable behavior. Piezoelectricity, however, exhibits a linear strain response to the electric field, an odd function that inverts with field reversal, and often shows hysteresis in ferroelectric materials due to irreversible domain wall motion under cycling fields. In practical applications, electrostriction supports symmetric, non-polarity-dependent actuation, eliminating the need for poling and enabling uniform deformation in both field directions, which is advantageous for devices requiring bidirectional or creep-free operation. Piezoelectric materials can deliver higher instantaneous strains but their polarity-sensitive response and potential complicate designs in high-cycle or unbiased environments, often necessitating careful field control.
AspectElectrostrictionPiezoelectricity
CoefficientQ (m⁴/C²), typically ~10⁻² for ceramics like d (pm/V), typically 100–750 for
Typical strain level0.01–0.2% at 1–3 kV/mm (e.g., : ~0.2%)0.1–0.7% at 1–3 kV/mm (e.g., : ~0.16%)
Units and scaleQuadratic coupling via Linear coupling via field

Versus Other Electro-Mechanical Effects

Electrostriction differs from flexoelectricity in that the latter involves the generation of due to a in materials lacking inherent , whereas electrostriction produces directly from an applied regardless of material . In flexoelectricity, deformation induces an electric response, often prominent in nanoscale structures or heterogeneous materials, as described in theoretical frameworks linking to . Conversely, electrostriction is a universal electro- where the drives symmetric distortion without requiring . The electrocaloric effect, which entails a reversible change in a material under an applied due to electrocaloric variations, contrasts with electrostriction's primary focus on mechanical deformation rather than thermal response. While both phenomena arise from interactions with polarizable materials, the electrocaloric effect is thermodynamic in , often exploited for cooling applications, and does not inherently produce macroscopic strain. Electrostriction, by comparison, manifests as dimensional changes without significant net shifts under isothermal conditions. In relation to converse magnetostriction, also known as the Villari effect, electrostriction is driven by inducing through electrostatic interactions, whereas converse magnetostriction results from altering and thus lattice dimensions in ferromagnetic materials. This magnetic analog shares a quadratic field dependence but operates via spin-orbit coupling rather than . Electrostriction's reliance makes it applicable in non-magnetic dielectrics, broadening its utility beyond magnetostrictive systems. Electrostriction stands out among these effects as a phenomenon always present in any dielectric material, producing proportional to the square of the without causing a net change in , unlike linear piezoelectric responses that it may complement in polar materials. This inherent universality and ensure electrostriction contributes to electro-mechanical behavior even in centrosymmetric crystals where is absent.

Materials and Properties

Common Materials

Electrostriction is a universal phenomenon observed in all dielectric materials, manifesting in common insulators such as glass, mica, and various polymers including poly(methyl methacrylate) (PMMA) and polystyrene. These materials typically exhibit low electrostrictive coefficients, resulting in minimal strains under applied electric fields that are often negligible for practical applications but confirm the effect's presence across everyday dielectrics. In ceramics like (BaTiO₃), electrostriction becomes observable in the paraelectric cubic phase above the , where the material lacks spontaneous polarization yet responds quadratically to . First-principles calculations for this phase yield electrostrictive coefficients of Q₁₁ = 0.115 m⁴/C², Q₁₂ = 0.033 m⁴/C², and Q₄₄ = 0.041 m⁴/C², aligning closely with experimental measurements and highlighting the effect's intrinsic nature in such high-permittivity ceramics despite the small overall strain. Early investigations demonstrated electrostriction in liquids and gases, including and dielectric oils, through measurable volume changes induced by , underscoring the effect's occurrence even in fluid . The extent of electrostriction in these common materials depends strongly on the , as higher permittivity amplifies the and thus the resulting for a given , while temperature influences the effect by altering permittivity and molecular mobility, particularly near relaxation frequencies or phase boundaries.

Enhanced or Giant Electrostriction Materials

Relaxor ferroelectrics, such as Pb(Mg<sub>1/3</sub>Nb<sub>2/3</sub>)O<sub>3</sub>-PbTiO<sub>3</sub> (PMN-PT), have been engineered to exhibit enhanced electrostrictive responses due to their disordered polar structures and proximity to ferroelectric transitions, achieving electrostrictive coefficients Q up to 0.1 m<sup>4</sup>/C<sup>2</sup>. These materials leverage nanoscale polar domains that amplify the quadratic electromechanical coupling, enabling large, low-hysteresis strains under moderate electric fields compared to conventional ferroelectrics. In polymer-based systems, terpolymers like poly(vinylidene fluoride-trifluoroethylene-chlorofluoroethylene) [P(VDF-TrFE-CFE)] demonstrate giant electrostriction through relaxor behavior induced by defect incorporation, yielding longitudinal strains exceeding 4% at electric fields around 100 MV/m. This high strain arises from field-induced polarization fluctuations in the polymer chains, combined with the material's flexibility and high permittivity, making it suitable for soft actuators. A notable breakthrough occurred in with the discovery of giant electrostriction in gadolinium-doped ceria (Ce<sub>0.8</sub>Gd<sub>0.2</sub>O<sub>1.9</sub>) thin films, where oxygen vacancy dynamics under generate stresses up to 500 MPa, corresponding to strains on the order of 0.1-1% depending on film thickness and . This stems from electroactive defect dipoles formed by aliovalent doping, which couple strongly to the , outperforming traditional ionic conductors in electromechanical efficiency. Research in the has focused on nanocomposites to further elevate Q values, such as reduced graphene oxide/polydimethylsiloxane (rGO/PDMS) hybrids that integrate high polarization coefficients with polymer compliance, achieving electrostrictive coefficients exceeding those of pure polymers by leveraging interfacial charge effects. These composites enhance overall through nanoscale fillers that promote uniform field distribution and defect . Recent advancements emphasize lead-free alternatives for giant electrostriction, such as doped oxides and relaxor ferroelectrics, to address environmental concerns and enhance performance in applications like medical devices, sensors, and .

Magnitude and Measurement

Typical Magnitudes

Electrostriction typically produces longitudinal strains ranging from $10^{-6} to $10^{-3} in various materials under applied of 1 to 10 kV/mm. These values are significantly smaller than the strains observed in piezoelectric materials, which commonly reach 0.1% to 1% under comparable fields. For example, in relaxor ferroelectrics such as Pb(Mg_{1/3}Nb_{2/3})O_3 (), electrostrictive strains of approximately $10^{-4} are typical under fields up to several kV/mm. The strain magnitude depends quadratically on the electric field strength, expressed as S \propto E^2, which distinguishes electrostriction from linear electro-mechanical effects. It also scales with the square of the material's , as higher enhances and thus the resulting deformation. influences the response, with strains often peaking near phase transitions where is maximized, such as the maximum dielectric temperature in relaxors. In cases of "giant" electrostriction, certain relaxor ferroelectric polymers can exhibit strains up to 4% at high fields of 75 to 200 MV/m, though this is limited by the risk of dielectric breakdown and material instability. For instance, electron-irradiated poly(vinylidene fluoride-trifluoroethylene) achieves around 4% strain at approximately 100 MV/m. Recent developments (as of 2024) in lead-free ceramics, such as sodium titanate-based thin films, have reported electrostrictive coefficients Q up to 0.32 m⁴/C² with strains ~0.2% at lower fields (~50 MV/m), and effective piezoelectric coefficients exceeding 10^4 pC/N in materials like gadolinium-doped ceria and lead halide perovskites. The electrostriction coefficient Q, which relates strain to the square of polarization (S = Q P^2), varies widely across material classes and governs the overall magnitude.
Material ClassTypical Q (m⁴/C²)Example MaterialReference
Relaxor ferroelectrics0.01–0.1BNT-based relaxors (0.030)
Relaxor ferroelectric polymers>10–40s-tetrapolymer (>40)
Low-permittivity dielectrics>1Polyurethanes (~10^5–10^6, inferred from M and \epsilon)

Measurement Techniques

Direct displacement measurements of electrostriction typically employ strain gauges or optical interferometry to quantify strain induced by applied electric fields, either direct current (DC) or alternating current (AC). Strain gauges, attached directly to the sample surface, detect minute changes in length or thickness under electric field excitation, providing a straightforward electrical readout of deformation. This method is particularly useful for bulk materials where surface bonding ensures reliable contact, though it requires calibration to account for gauge sensitivity and field uniformity. Interferometry, on the other hand, offers non-contact, high-resolution detection of sub-nanometer displacements by analyzing interference patterns from laser light reflected off the sample. To isolate the quadratic electrostrictive response from linear piezoelectric effects, especially in materials exhibiting both, double-beam interferometry is widely adopted. In this setup, two beams probe the sample's front and rear surfaces simultaneously, enabling differential measurement that cancels out bending or linear contributions while amplifying the field-squared dependence characteristic of electrostriction. Configurations such as Mach-Zehnder or common-path interferometers enhance stability against environmental vibrations, achieving resolutions down to picometers for strains under high-frequency fields up to several MHz. These techniques are essential for thin films and ceramics, where can dominate at low fields, and have been validated in separating pure electrostriction in non-piezoelectric dielectrics. Impedance spectroscopy provides an indirect approach to infer the electrostriction coefficient Q by analyzing the between dielectric and under bias fields. The method involves applying a small AC signal superimposed on a and measuring the complex impedance spectrum to identify shifts in frequencies, which arise from electrostrictive stiffening or softening of the material. From these spectra, Q is extracted using models that relate changes to via the electrostrictive tensor, particularly effective for resonant modes in the kHz to GHz range. This non-destructive technique is advantageous for in-situ characterization of multilayer devices, though it relies on accurate modeling to disentangle contributions from and . Measuring electrostriction presents several challenges, including the need for high —often exceeding 10 kV/mm—to produce observable strains in low-permittivity materials, which risks dielectric breakdown or electrode degradation. Separating electrostrictive effects from requires AC excitation at frequencies above diffusion rates, as low-frequency measurements can confound Joule heating-induced expansions with field-driven strains. In materials with residual piezoelectricity, such as relaxor ferroelectrics, distinguishing the quadratic response demands precise field polarity reversal or multi-beam setups to suppress linear artifacts. These issues necessitate controlled environments, like chambers to minimize effects, and advanced to enhance signal-to-noise ratios.

Applications

Actuators and Transducers

Electrostriction is widely employed in actuators and transducers due to its ability to produce symmetric responses proportional to the square of the applied , enabling precise and repeatable motion without the need for material poling. Unlike piezoelectric effects, which require non-centrosymmetric and exhibit , electrostriction operates effectively in centrosymmetric materials, offering hysteresis-free performance under unbiased conditions and enhanced resistance for long-term applications. This makes it particularly suitable for devices demanding high precision and reliability, such as those in and acoustics. In precision actuators for optical systems, electrostrictive materials enable fine mirror positioning and shape control, critical for maintaining focus in . For instance, multilayer electrostrictive ceramics based on lead magnesium niobate (PMN) have been integrated into deformable mirrors for the (JWST), where they adjust surface contours to compensate for thermal distortions with sub-micron accuracy. Polymer films, such as those derived from electrostrictive graft elastomers, further extend these applications by providing lightweight, flexible actuation, achieving strains up to 0.1% under moderate fields without remanent deformation. Ultrasonic transducers utilizing electrostriction excel in due to their high-frequency response and lack of , allowing for clear, artifact-free signal generation and reception. Relaxor ferroelectric ceramics like PMN-PT are commonly used in these devices, where a enhances the effective electromechanical coupling while preserving the quadratic strain-field relationship for operation up to several MHz. Clinical evaluations have shown that electrostrictive transducers match the performance of traditional PZT-based ones in pulse-echo , with added benefits in reduced mismatch for better tissue penetration. For vibration control in structural applications, electrostrictive relaxor ceramics provide effective through active strain modulation, suppressing resonances in and components. These materials, often configured in stack or unimorph geometries, generate counteracting forces proportional to the field squared, enabling real-time adaptation to dynamic loads without the aging effects seen in poled piezoelectrics. In particular, PMN-based ceramics have demonstrated significant reduction in structures, with densities supporting low-frequency actuation below 1 kHz.

Sensors and Emerging Uses

Electrostrictive nanocomposites, such as those incorporating reduced (rGO) with (PDMS), enable high-sensitivity and sensors suitable for wearable devices due to their flexibility and enhanced properties. These materials leverage giant electrostriction to detect subtle mechanical deformations, providing improved responsiveness in applications like health monitoring wearables, where sensitivities can reach levels comparable to traditional piezoelectric sensors but with added conformability to . Electrostrictive polymers are employed in energy harvesting devices that convert mechanical vibrations into through quadratic coupling, offering advantages in low-frequency ambient sources like human motion or machinery. For instance, pre-stretched electrostrictive films coupled with nonlinear biasing techniques have demonstrated efficient harvesting from vibrations, achieving power densities suitable for powering small sensors without external batteries. This approach exploits the material's intrinsic electrostriction to generate voltage under oscillatory strain, enhancing self-powered systems in remote or wearable contexts. In 2024, advances in bio-compatible electrostrictors, such as oxygen-defective ceria-based thin films, have enabled soft robotics actuators with enhanced electromechanical performance for biomedical applications, exhibiting large strains under low voltages while maintaining compatibility with biological tissues. Concurrently, integration of electrostrictive materials into micro-electro-mechanical systems (MEMS) supports Internet of Things (IoT) devices, where compact harvesters and sensors utilize the effect for vibration-based energy conversion and precise deformation detection in networked environments. The hysteresis-free nature of electrostriction positions it as a promising mechanism for tunable deformation in antennas, allowing precise, reversible adjustments to radiation patterns without residual strain artifacts common in other electroactive materials.

References

  1. [1]
    Defining “giant” electrostriction - AIP Publishing
    May 5, 2022 · Electrostriction is an electromechanical phenomenon that exists in all dielectrics. In the direct effect, the induced strain or stress is ...
  2. [2]
    Electrostriction - an overview | ScienceDirect Topics
    Electrostriction is the ability of all insulator or dielectric materials to develop mechanical deformation if subjected to an electric field.Missing: reliable | Show results with:reliable
  3. [3]
    Data-driven methods for discovery of next-generation electrostrictive ...
    Dec 8, 2022 · All dielectrics exhibit electrostriction, i.e., display a quadratic strain response to an electric field compared to the linear strain ...
  4. [4]
    Electrostrictive effect in ferroelectrics: An alternative approach to ...
    Jan 15, 2014 · Electrostriction plays an important role in the electromechanical behavior of ferroelectrics and describes a phenomenon in dielectrics where the strain varies ...
  5. [5]
    Electrostriction - an overview | ScienceDirect Topics
    Electrostriction is defined as the shape change that occurs in any dielectric material under the application of an external electric field, exhibiting a ...Missing: reliable | Show results with:reliable
  6. [6]
  7. [7]
    (PDF) Foundations of Piezoelectrics - ResearchGate
    Apr 2, 2023 · Several expressions for the electrostrictive coefficient Q have been proposed so far. From the data obtained by independent experimental methods ...
  8. [8]
  9. [9]
  10. [10]
  11. [11]
  12. [12]
    Electrostriction
    Electrostriction is an interaction in solid dielectric materials, where an electric field applied on the material generates the deformation of the material ...
  13. [13]
    [PDF] Comparison of Piezoelectric, Magnetostrictive, and Electrostrictive ...
    The phenomenon of electrostriction, however, is fundamentally different from the converse piezoelectric effect. The unit cell in an electrostrictive material ...
  14. [14]
    4: Review of Electrostrictive Materials - Books
    Apr 22, 2020 · Electrostriction is defined as the electromechanical coupling in all electrical-nonconductor (dielectric, insulator) materials. Under the ...
  15. [15]
    None
    ### Key Equations for Electrostriction
  16. [16]
    Physical & Piezoelectric Properties of Products | APC Int.
    Piezoelectric Properties ; Piezoelectric Charge Constant (10-12 C/N or 10-12 m/V) ; d · 300, 300 ; Piezoelectric Voltage Constant (10-3 Vm/N or 10-3 m2/C).
  17. [17]
    Dopant Concentration Controls Quasi-Static Electrostrictive Strain ...
    ... electrostriction strain coefficient with magnitude on the order of 10–18 m2/V2. The quasi-static (f < 1 Hz) electrostriction strain coefficient for undoped ...
  18. [18]
    [PDF] Lattice, elastic, polarization, and electrostrictive properties of BaTiO3 ...
    Our calculated electrostrictive coefficients of cubic BaTiO3, compared with the experimental data from given references. Q11. (m4 C−2). Q12. (m4 C−2). Q44. (m4 ...
  19. [19]
    Electrostriction in Aqueous Solutions of Electrolytes - AIP Publishing
    The change in solvent volume near ions can be related, using an appropriate molecular model, to the apparent change in the volume of the salt upon dissolution.
  20. [20]
    Ultralow Hysteresis and a Giant Electrostrictive Coefficient in Pb(Mg ...
    May 20, 2024 · Pb(Mg1/3Nb2/3)O3 (PMN) is a typical electrostrictive material. ... Although Q33 is commonly used to evaluate the electrostrictive property of ...
  21. [21]
    Ultrahigh electrostrictive strain and its response to mechanical ...
    Mar 1, 2024 · Of particular significance is that a high electrostrictive coefficients M33 = 12.2 × 10−16 m2/V2, together with a high mechanical work density ...
  22. [22]
    Structure-property relationships in three-phase relaxor-ferroelectric ...
    P(VDF-TrFE-CFE) terpolymers show the highest electrostrictive strain among all RF polymers known so far [Citation18, Citation20]: about 7% in the thickness ...
  23. [23]
    Core-free rolled actuators for Braille displays using P(VDF-TrFE-CFE)
    Experimental testing of 17 actuators demonstrates significant strains (up to 4%) and blocking forces (1 N) at moderate electric fields (100 MV m-1). A novel ...
  24. [24]
    Giant electrostriction in Gd-doped ceria - PubMed
    Nov 14, 2012 · Gd-doped CeO(2) exhibits an anomalously large electrostriction effect generating stress that can reach 500 MPa.Missing: thin films
  25. [25]
    In-situ extended X-ray absorption fine structure study of ...
    Jan 29, 2015 · Recently, giant electrostriction, producing stress on the order of hundreds of MPa, was reported in thin films of 20 at. % Gd-doped ceria (⁠ ...
  26. [26]
    Fluoropolymer ferroelectrics: Multifunctional platform for polar ...
    May 12, 2023 · Certain fluoropolymers have ferroelectric properties and attractive mechanical properties for a wide range of applications.
  27. [27]
    Development and analysis of flexible high-performance ...
    Aug 10, 2025 · Here, we provide a sustainable, innovative, flexible, adaptable, and environmentally stable P(VDF-TrFE)/rGO-NdMnO3 hybrid polymer nanocomposite- ...
  28. [28]
    Flexible piezoelectric energy harvester based on self-poled ...
    These composites combine the high piezoelectric performance of the fillers with the flexibility of the polymer matrix. Examples include organic/inorganic ...<|control11|><|separator|>
  29. [29]
    [PDF] Electrostrictive Strain/Dielectric Properties of Relaxor Ferroelectrics
    A wide temperature range has been investigated for dielectric permittivity,. K(T), strain-E field and polarization-E field response. Strain levels achieved for ...<|control11|><|separator|>
  30. [30]
    Electrostrictive Strain in Low-Permittivity Dielectrics | Request PDF
    Aug 5, 2025 · The magnitude of the displacements varies between 10–2 to 10–3 under 1 MV/m electric field. By comparison, the field-induced displacements in ...
  31. [31]
    [PDF] Large electrostrictive response in lead halide perovskites
    Electrostriction is an electromechanical coupling with strain proportional to the square of applied electric field, while piezoelectric strain is linear with ...Missing: differences | Show results with:differences
  32. [32]
    [PDF] Electrostriction in field-structured composites - OSTI
    Dec 14, 1998 · When a suspension of dielectric particles is exposed to a uniaxial electric field, the induced dipole moments will cause the particles to chain ...
  33. [33]
    Temperature dependence of high field electromechanical coupling ...
    Apr 8, 2010 · Electrostriction coefficients are shown to be temperature dependent in these ceramic materials, with different values above and below the ...
  34. [34]
    [PDF] Micromechanics of ferroelectric polymer-based electrostrictive ...
    Abstract. Electrostriction refers to the strain induced in a dielectric by electric polarization, which is usually very small for practical application.
  35. [35]
    [PDF] Electrostrictive Graft-Elastomers (G-Elastomers)
    Aug 26, 2019 · Physically cross-linked (formed crystal domains) elastomers that offer field-induced strain by primarily electrostriction. S train, (%).
  36. [36]
    Giant room-temperature electrostrictive coefficients in lead-free ...
    Jan 16, 2018 · In summary, a large room-temperature electrostrictive coefficient of 0.030 m4/C2 is attained in BNT-based relaxor ferroelectrics, as the direct ...
  37. [37]
    Relaxor ferroelectric polymer exhibits ultrahigh electromechanical ...
    Mar 24, 2022 · We found high EM performance of P(VDF-TrFE-CFE-FA) s-tetrapolymers, with a k33 of 88% and d33 of –1050 pm/V even at low electric fields (<50 ...
  38. [38]
    [PDF] Material parameters for electrostriction
    Electrostriction is often described by a phenomenological tensor relating a material's deformation to an applied electric field. However, this tensor is not ...
  39. [39]
    Electrostriction measurements on low permittivity dielectric materials
    Q coefficients for the samples from both techniques are given in Table 1. It may be seen in the tabulation that all results agree well. This helps us achieve ...
  40. [40]
    Electrostriction: Nonlinear Electromechanical Coupling in Solid ...
    Electrostriction is the basis of electromechanical coupling in all insulators. The quadratic electrostrictive strain x ij associated with induced polarization ...Missing: sources | Show results with:sources
  41. [41]
    A double-beam common path laser interferometer for the ...
    We report in this paper the development of a modified Mach-Zehnder-type double-beam common path laser interferometer for the measurement of electric ...
  42. [42]
    Challenges in double-beam laser interferometry measurements of ...
    Jun 2, 2022 · This work serves to identify these challenges by demonstrating d 33,f measurements of fully released PZT films using a commercial double-beam laser ...Missing: method electrostriction
  43. [43]
    (PDF) Electrostrictive materials: Characterization and applications ...
    Aug 9, 2025 · The resonance equations for the DC biased length extensional resonator are presented and it is shown that DC biased resonance techniques can be ...
  44. [44]
    Electrostrictive resonances in (Ba0.7Sr0.3)TiO3 thin filmsat ...
    Jul 26, 2004 · The dielectric spectra of the films exhibit resonance phenomena in the 4 – 7 GHz range when a bias field is applied. It is shown that the main ...Missing: shift | Show results with:shift
  45. [45]
    [PDF] Interferometric investigation of quadratic electrostriction in solid ...
    Jan 1, 1996 · In interferometric measurements of quadratic electrostriction, anomalies in the frequency-dependence of the recorded strain signa! are sometimes ...
  46. [46]
    Electrostrictive actuators and their use in optical applications
    Precision gets greater in the machining industry, lines get finer in semiconductor processing, tracks get closer in information.Missing: telescopes | Show results with:telescopes
  47. [47]
    Electrostrictive ceramic actuators to shape mirror of next space ...
    Apr 28, 2010 · Electrostrictive ceramic actuators to shape mirror of next space telescope. According to a NASA Tech Brief, the Next Generation Space Telescope ...
  48. [48]
    [PDF] electrostrictive graft elastomers and applications
    Materials that sustain mechanical displacement under controlled electrical excitation are needed as actuators for many applications. For aerospace applications,.
  49. [49]
    DC-biased electrostrictive materials and transducers for medical ...
    This paper reviews the most important material properties and summarizes the advantages of using electrostrictive materials in medical imaging applications. By ...
  50. [50]
    Electrostrictive Transducers for Medical Ultrasonic Applications
    Clinical imaging results indicate that electrostrictive transducers perform as well as conventional PZT-type transducers with the additional benefit of ...
  51. [51]
    (PDF) Modeling and characterization of electrostrictive ceramics
    Aug 9, 2025 · The model allows for the non-zero piezoelectric behavior found in some nominally electrostrictive materials. The quasistatic non-linear ...<|separator|>
  52. [52]
    [PDF] Electrostrictive Ceramics for Low-Frequency Active Transducers
    Abstract—Electrostrictive ceramics based on Pb(Mg1=3. Nb2=3)O3 have demonstrated promise as low-frequency ac- tive materials. They broaden the range of ...<|separator|>
  53. [53]
    Electrostrictive polymers for mechanical energy harvesting
    Feb 6, 2012 · Electrostriction is generally defined as a quadratic coupling between strain (Sij) and polarization (Pm):8, 9. equation image (1). where ...
  54. [54]
    Electrostrictive polymer harvesting using a nonlinear approach
    Electrostrictive polymers are a new type of material for high-efficiency, low-power energy harvesting from ambient vibrations. Owing to a decrease of power ...
  55. [55]
    Mechanical characterization of an electrostrictive polymer for ...
    Jun 28, 2012 · ... harvesting energy from vibration sources, such as human motion.10,11. Electrostrictive polymers exhibit very attractive characteristics, and ...
  56. [56]
    Oxygen-defective electrostrictors for soft electromechanics - PMC - NIH
    Aug 30, 2024 · This study explores the potential of ceria-based thin films as electromechanical actuators for flexible electronics. Oxygen-deficient fluorites, ...