Fact-checked by Grok 2 weeks ago

Negative thermal expansion

Negative thermal expansion (NTE) is a counterintuitive in which certain materials contract in volume or linear dimensions upon heating under constant , in direct opposition to the positive displayed by most substances. This behavior is quantified by a negative linear of (α_L), often ranging from -1 to -100 K⁻¹, where the material's size decreases as temperature increases. The mechanisms driving NTE can be broadly classified into three categories: flexible network structures, where low-frequency phonon modes or transverse vibrations in open frameworks cause net contraction; atomic radius contraction, involving charge transfer or electronic effects that reduce interatomic distances; and magnetovolume effects, where changes in magnetic ordering lead to volume reduction. Notable examples include zirconium tungstate (ZrW₂O₈), which exhibits isotropic NTE of approximately -9 ppm K⁻¹ from 0.3 K to 1050 K due to rigid unit modes in its framework; scandium fluoride (ScF₃), showing giant NTE from transverse vibrations; and β-eucryptite, a lithium aluminosilicate with uniaxial NTE along its c-axis. Historically, low or near-zero thermal expansion was first systematically studied in the iron-nickel alloy Invar (Fe-36Ni) by Charles Édouard Guillaume in 1897, for which he received the Nobel Prize in Physics in 1920, though true negative expansion in a wider range of materials was identified later, with ZrW₂O₈ marking a breakthrough in 1996. NTE materials hold significant promise for applications requiring precise control of thermal expansion, such as in precision , circuits, components, and composites where they compensate for the expansion of other materials to achieve zero or tailored overall expansion. Advances as of the include the development of giant NTE materials with coefficients exceeding -30 ppm K⁻¹, like certain nitrides (e.g., Mn₃Zn₀.₅Sn₀.₅N), and microstructural in metamaterials to enhance functionality across broader temperature ranges and directions. More recent progress as of 2025 features ultrastrong NTE in compositionally complex alloys and tunable expansion in structures, alongside integration of NTE materials in solid oxide fuel cells to reduce .

Fundamentals

Definition and Phenomenon

Negative thermal expansion (NTE) refers to the unusual phenomenon in which a material's dimensions decrease as its temperature increases under . This counterintuitive behavior is characterized by a negative of , denoted as α < 0, where the linear is defined by the formula α = (1/L)(dL/d), with L representing the material's length and T the . In typical materials, positive thermal expansion predominates due to anharmonic lattice vibrations, which cause atoms to oscillate with greater amplitude in asymmetric potential wells, effectively increasing the average interatomic distances upon heating. NTE defies this expectation, leading to contraction instead of expansion. The effect can be observed in linear, areal, or volumetric dimensions, depending on the material's symmetry. In anisotropic materials, NTE often appears transversely—perpendicular to certain atomic bonds—while longitudinal expansion may occur along the bond directions, resulting from the excitation of low-energy transverse vibrational modes at lower temperatures compared to longitudinal ones. Thermodynamically, NTE complies with the second law and emerges from specific lattice dynamics, such as those driven by phonon modes that favor contraction over expansion.

Historical Discovery

The phenomenon of negative thermal expansion (NTE) in solids was first observed in the early 20th century, with Karl Scheel reporting shrinkage upon heating in quartz and vitreous silica at low temperatures in 1907. These early findings highlighted anomalous contraction behaviors but were limited to specific temperature ranges and materials, without broader systematic investigation. In the 1950s, studies on lithium aluminum silicates (LAS) revealed NTE properties, marking initial recognition of the effect in crystalline frameworks, though it remained sporadic and not fully characterized. Systematic research accelerated in the 1990s, driven by precise diffraction measurements that confirmed isotropic NTE over wide temperature ranges. A pivotal contribution came from Arthur W. Sleight and colleagues, who in 1995 identified NTE in cubic , followed by their 1996 discovery of exceptional NTE in , contracting from 0.3 K to its decomposition temperature near 1050 K with a coefficient of approximately -9 × 10⁻⁶ K⁻¹. This work by Sleight's group emphasized open-framework structures as key enablers, shifting focus from isolated anomalies to engineered materials. Concurrently, the terminology evolved from "anomalous expansion" to in scientific literature, reflecting the growing acceptance of NTE as a distinct, tunable property rather than a mere outlier. In the 2000s, advancements in neutron scattering techniques unveiled the underlying phonon modes responsible for NTE, as demonstrated in a 2001 high-pressure inelastic neutron scattering study on , which linked transverse vibrations of oxygen atoms to the contraction mechanism. Post-2000 milestones expanded NTE to diverse systems, including the confirmation of significant NTE in phases around 2002 and the discovery of NTE in (MOFs) during the 2000s, such as the 2008 report of α ≈ -4 × 10⁻⁶ K⁻¹ in . By the 2020s, research has increasingly targeted room-temperature NTE materials, with developments in oxygen-redox active compounds exhibiting α = -14.4 × 10⁻⁶ K⁻¹ and extended NTE in PbTiO₃-based perovskites over broad temperature ranges (as of 2025), enabling practical applications through controlled synthesis.

Mechanisms

General Origins

Negative thermal expansion (NTE) fundamentally arises from specific vibrational modes in the lattice, particularly low-frequency phonon modes characterized by negative . The for a phonon mode, defined as \gamma = -\frac{d \ln \omega}{d \ln V}, quantifies the coupling between the mode's frequency \omega and the crystal volume V; when \gamma < 0, an increase in temperature excites these modes, which respond by increasing their frequency under compression, effectively generating tensile stress that contracts the lattice. This phonon-driven mechanism dominates NTE across diverse materials, as confirmed by and density-of-states measurements that highlight the prominence of such modes with large negative \gamma values, often separated by a from higher-frequency vibrations. Anharmonic effects play a crucial role in enabling these negative \gamma values, as the non-linear nature of interatomic potentials allows for asymmetric vibrational responses. In particular, transverse vibrations of rigid structural units, such as polyhedra in framework materials, lead to contraction because the anharmonic potential pulls atoms closer together as vibrational amplitudes increase with temperature. The contribution of a single mode to the volumetric thermal expansion coefficient \alpha_V can be approximated in the quasi-harmonic framework as \alpha_{V,i} = \frac{\gamma_i c_i}{B V}, where c_i is the mode's heat capacity contribution, B is the bulk modulus, and V is the volume; for modes with \gamma_i < 0, this yields a negative term that reduces or reverses overall expansion when dominant. This simplified relation underscores how low-energy, anharmonically coupled phonons drive NTE by prioritizing lattice tension over the usual dilatational effects of thermal motion. Structurally, NTE requires architectures that permit such modes, typically open lattices or flexible frameworks where atoms or units can undergo concerted rotations or tilts without significant bond stretching. These configurations contrast with rigid close-packed systems, where high coordination and dense packing favor positive Grüneisen parameters and conventional expansion, limiting the space for transverse or librational motions essential for contraction. At low temperatures, quantum mechanical effects further contribute to NTE through zero-point motion, where the ground-state vibrational energy gradients across the lattice induce contraction as the zero-point amplitude effectively mimics thermal excitation but with quantum coherence. In certain cases, quantum tunneling between potential minima enhances this by allowing barrier penetration in anharmonic wells, amplifying the negative expansion observed below cryogenic temperatures.

In Close-Packed Structures

In close-packed structures, negative thermal expansion (NTE) arises primarily from the interplay between geometric constraints and vibrational dynamics, where specific phonon modes contribute negatively to the overall expansion coefficient. Unlike open-framework materials, densely packed atomic arrangements—such as face-centered cubic (fcc) or hexagonal close-packed (hcp) lattices—typically exhibit positive thermal expansion due to dominant anharmonic bond-stretching effects. However, NTE can emerge when certain low-frequency phonon modes possess negative Grüneisen parameters (γ < 0), indicating that their frequencies increase with volume expansion, leading to a net contraction upon heating. These modes often involve transverse vibrations that effectively shorten interatomic distances despite increased thermal energy. Geometric frustration plays a key role in such systems, particularly in structures where the dense packing of polyhedra or rigid units restricts thermal motion, favoring rotational instabilities over linear expansion. A model for this rotational NTE describes the apparent reduction in bond length due to angular fluctuations: the difference between the mean-square distances in a rotating unit approximates R(1 - (1/2)<θ²>T), where R is the unit radius, θ is the rotational angle, and T is temperature, highlighting how thermal excitation of librational modes shortens effective distances. Phonon mode analysis reveals that saddle-point-like modes in close-packed lattices are particularly influential, as their potential energy surfaces feature minima that promote inward atomic displacements upon excitation. These modes, often transverse optic phonons, yield negative contributions to volume expansion, approximated by the sum over relevant modes of γ_i (ℏω_i / k_B T), where γ_i is the mode , ℏω_i is the energy, k_B is Boltzmann's constant, and T is ; negative γ_i for these modes dominates at low temperatures, causing ΔV/V < 0. In close-packed ionic like rock-salt structured RbI, such modes lead to NTE below approximately 8 K. Examples of NTE in metals and alloys further illustrate these effects, often amplified by electronic contributions near the . In compounds such as InBi, which adopts a layered close-packed structure, pronounced NTE (α ≈ -85 × 10^{-6} K^{-1} parallel to layers at ) stems from anisotropic electron- coupling and instabilities that enhance negative Grüneisen contributions. Similarly, in heavy like UPt₃ (hexagonal close-packed), NTE at low temperatures arises from a two-component Fermi-liquid model, where magnetic and electronic fluctuations near the couple to lattice vibrations, yielding contraction coefficients up to -10 × 10^{-6} K^{-1} below 20 K. These electronic effects modulate spectra, making γ more negative in dense metallic environments. NTE in close-packed systems is typically confined to low temperatures, where quantum effects and minimal allow negative mode contributions to prevail; it diminishes at higher temperatures as higher-order anharmonic terms restore positive expansion, often transitioning to near-zero or positive values above 100 in metals like Zn (negative perpendicular expansion below 10 ). This temperature dependence underscores the delicate balance in dense lattices, where vibrational is overcome by at elevated T.

Exotic Mechanisms

Electronic contributions to negative thermal expansion (NTE) arise from thermal alterations in electronic structure, such as charge density waves (CDWs) or valence transitions, particularly in transition metal compounds where Fermi surface instabilities play a role. In these systems, heating can trigger electronic rearrangements that reduce lattice volume, distinct from dominant vibrational effects. For instance, in the layered compound YbMn₂Ge₂, a dual mechanism involving CDW formation and Yb valence transition from ~2.40 to ~2.82 induces NTE with a volumetric coefficient of α_v = -32.9 × 10⁻⁶ K⁻¹ over 400–575 K, driven by magnetovolume effects tied to Mn electronic instabilities near the Néel temperature (~510 K). Similarly, in monolayer 1T-NbSe₂, CDW transitions distort the "stars of David" lattice motifs, leading to pronounced NTE through electronic modulation of bond lengths and angles. Magnetic effects contribute to NTE via magnetovolume coupling, where changes in magnetic ordering alter lattice parameters. In rare-earth compounds, antiferromagnetic (AFM) transitions often induce contraction as the system shifts to a paramagnetic state with reduced volume . This arises from exchange interactions and spin fluctuations, quantified by spontaneous volume magnetostriction ω_s ∝ M² + ξ², where M is and ξ represents fluctuations. Exemplary cases include R₂Fe₁₇ (R = rare earth) intermetallics, exhibiting NTE during ferrimagnetic-to-paramagnetic transitions due to AFM-like volume reduction, and antiperovskite Mn₃Cu_{1-x}Ge_xN (x ≈ 0.5), where gradual AFM moment development near yields giant NTE from magnetovolume effects. Recent examples include giant NTE in PrMnO₃, with coefficients exceeding -100 ppm K⁻¹ over a 1000 K range, driven by successive magnetic phase transitions and magnetovolume coupling. Topological and quantum mechanisms enable NTE in low-dimensional systems through non-trivial vibrational or spin degrees of freedom. In graphene-like 2D materials, flexural (out-of-plane bending) modes dominate, as their anharmonic coupling with in-plane stretching favors lattice contraction upon heating; the thermal expansion coefficient remains negative up to ~1000 K, with α ≈ -7 × 10⁻⁶ K⁻¹ at room temperature, arising from increased transverse fluctuations that effectively shorten projected bond lengths. Theoretical predictions post-2020 extend this to quantum spin liquids (QSLs) in frustrated magnets, where emergent gauge fields and fractionalized excitations couple to phonons, inducing NTE; for example, in the frustrated spinel CdCr₂O₄, a band of localized magnetic excitations in the half-magnetization plateau phase (above 27 T, 4.2–10.4 K) drives NTE via strong spin-lattice coupling, analogous to QSL dynamics. Hybrid cases, such as photoinduced NTE, link electronic and optical responses in perovskites through light-matter interactions that transiently alter structure.

Materials

Inorganic Crystals

One of the classic examples of negative thermal expansion (NTE) in inorganic crystals is zirconium tungstate (ZrW₂O₈), which exhibits isotropic NTE with a linear coefficient α ≈ -9 × 10⁻⁶ K⁻¹ over a broad temperature range from 0.3 to 1050 K. This material adopts a cubic structure (space group P2₁3) composed of corner-sharing ZrO₆ octahedra and WO₄ tetrahedra, where the NTE arises primarily from transverse thermal vibrations of oxygen atoms that induce coupled rotations and tilting of the polyhedra, effectively contracting the lattice upon heating. The NTE persists up to the material's decomposition temperature near 1050 K, though the coefficient decreases to approximately -5 × 10⁻⁶ K⁻¹ above 450 K due to phase transitions from the low-temperature α-phase to the β-phase. Other crystals display similar but temperature-dependent NTE behaviors influenced by structural transitions. molybdate (HfMo₂O₈) features a cubic structure analogous to ZrW₂O₈ and exhibits NTE with a linear α ≈ -4 × 10⁻⁶ K⁻¹ around , driven by low-frequency optic modes causing polyhedral rotations; it remains stable up to high temperatures. pyrovanadate (ZrV₂O₇) shows NTE only in its high-temperature cubic above approximately 375 K up to 1075 K, while lower temperatures exhibit positive expansion due to transitions to ordered structures that restrict the quasi-rigid unit modes responsible for contraction. In crystals, NTE often manifests uniaxially in -like structures or isotropically in types. Materials with ZrF₄-like architectures, such as (UF₄), display intrinsic uniaxial NTE below along the chain direction, driven by anisotropic dynamics and softening, with the monoclinic enabling contraction via fluorine atom displacements. In contrast, scandium trifluoride (ScF₃) exhibits isotropic NTE with α ≈ -10 × 10⁻⁶ K⁻¹ over an exceptionally wide range from 10 K to 1100 K in its cubic ReO₃-type , attributed to transverse vibrations of fluorine atoms in the corner-sharing ScF₆ octahedra that promote rigid unit mode rotations without significant . Recent advancements post-2015 have identified NTE in bismuth-based perovskites and certain cyanides, though synthesis remains challenging. In PbTiO₃-type perovskites like Bi₀.₆Na₀.₄VO₃, NTE occurs during the tetragonal-to-cubic phase transition, resulting in volume shrinkage linked to charge transfer and octahedral tilting; high-pressure (8 GPa) and high-temperature (1473 K) synthesis is required, but excess bismuth content leads to secondary phases like Bi₄V₂O₁₀, complicating phase purity. For cyanides, potassium cadmium dicyanoargentate (KCd[Ag(CN)₂]₃) shows NTE with α ≈ -15 × 10⁻⁶ K⁻¹ from 100 to 400 K in its framework structure, arising from low-energy bending modes of the linear [Ag(CN)₂]⁻ units; high-pressure studies confirm enhanced contraction under compression, but scalability is limited by sensitivity to moisture and synthetic complexity. Antiperovskite nitrides, such as Mn₃Zn₀.₅Sn₀.₅N, exhibit giant isotropic NTE with α exceeding -30 × 10⁻⁶ K⁻¹ over broad temperature ranges due to magnetovolume effects.

Frameworks and Composites

Metal-organic frameworks (MOFs) are porous materials that often display negative thermal expansion (NTE) due to flexible linker rotations and node distortions within their open architectures. In zeolitic imidazolate framework-8 (ZIF-8), a zinc-based MOF with linkers, thermal excitation leads to rotational motions of the linkers, contributing to NTE behavior, particularly in mixed-metal analogs where compositional inhomogeneity enhances negative expansion modes. The linear (α) for ZIF-8 typically ranges from positive values around +7 × 10^{-6} K^{-1} to tunable negative values in doped variants, influenced by guest molecule adsorption that modulates framework flexibility. Similarly, UiO-66, a zirconium-based MOF with terephthalate linkers, exhibits NTE through cooperative distortions of its Zr_6O_4(OH)_4 nodes, resulting in isotropic contraction upon heating. Defect engineering in UiO-66(Hf), a analog, amplifies this effect to colossal levels, with α ≈ -89 × 10^{-6} K^{-1} over 100–350 K, far exceeding typical MOF NTE. Guest-dependent tuning in these frameworks allows control of α from -10 to -50 × 10^{-6} K^{-1}, enabling applications in responsive materials. Zeolites and phosphate frameworks, as rigid open structures, demonstrate NTE primarily through transverse vibrations of polyhedral units that couple to reduce lattice dimensions. In aluminophosphate (AlPO_4) frameworks like AlPO_4-17, which adopts a hexagonal topology, low-frequency transverse modes of AlO_4 and PO_4 tetrahedra drive strong isotropic NTE, with an average α ≈ -11.7 × 10^{-6} K^{-1} from 18–300 K. Insertion of guest molecules, such as oxygen, further tunes this behavior by altering vibrational contributions, shifting the principal NTE direction. Zeolites, including germanosilicate ITQ-7 and ITQ-9, exhibit widespread NTE over broad temperature ranges due to similar rigid unit vibrations in their microporous cages, with α values as low as -9 × 10^{-6} K^{-1} in hydrated forms like HZSM-5. These materials highlight the role of framework openness in facilitating vibrationally driven contraction. Hybrid composites combining inorganic NTE phases with offer enhanced and tailorable expansion properties for practical use. Polymer-inorganic blends, such as or matrices filled with NTE ceramics like ZrW_2O_8 or β-eucryptite, achieve reduced or negative overall by leveraging the fillers' to counteract polymer dilation, with effective α down to -5 × 10^{-6} K^{-1} in optimized ratios. These systems provide mechanical flexibility absent in pure frameworks, making them suitable for thermal stresses in composites. Amorphous materials, including certain oxide , can exhibit NTE via rigid unit modes (RUMs) where polyhedral rotations mimic crystalline transverse vibrations without long-range order. In ZrO_2-based composites or related amorphous oxides, RUM-like dynamics contribute to low or negative expansion, though pure ZrO_2 glass typically shows positive behavior; blending with NTE phases like ZrW_2O_8 yields amorphous-leaning hybrids with α ≈ -9 × 10^{-6} K^{-1}. Post-2020 advancements include 3D-printed NTE composites, such as hyperbolically oriented metamaterials, which achieve tunable linear NTE with α ≈ -7.5 × 10^{-6} K^{-1} through architectural design, enabling customizable structures via additive . Tunability of NTE in these systems is achieved through doping, , or guest intercalation, expanding the temperature range and magnitude. Prussian blue analogs, cyanide-bridged frameworks like FeFe(CN)_6, display NTE (α ≈ -4 × 10^{-6} K^{-1}) from low-energy bending modes of the metal-cyanide links, tunable from negative to positive via redox intercalation of ions like Na^+ or water molecules. Doping with transition metals or applying alters the framework rigidity, extending NTE over 100–400 K in variants like M^{II}_2[M^{IV}(CN)_8], with α shifting by up to 20 × 10^{-6} K^{-1}.

Characterization and Theory

Measurement Techniques

Dilatometry and are fundamental techniques for directly measuring linear coefficients in materials exhibiting negative thermal expansion (NTE). In dilatometry, a push-rod mechanism contacts the sample ends to track dimensional changes as temperature varies, often using a (LVDT) for displacement detection. This method provides reliable data for bulk samples over wide temperature ranges, with resolutions achieving 10^{-8} K^{-1} in advanced setups. , particularly laser-based variants like the , employs optical interference patterns to monitor sub-micrometer length variations, offering superior precision for low-expansion materials without physical contact. These optical approaches mitigate mechanical artifacts but require controlled environments to avoid vibrations and thermal gradients, limiting their use in anisotropic samples where directional measurements are essential. X-ray and enable in-situ probing of atomic-scale parameters to quantify NTE through structural with . In , powder or single-crystal samples are subjected to variable-temperature scans, where shifts in positions reveal changes in interplanar spacing, allowing calculation of the linear coefficient α from Δa/a versus ΔT. complements this by providing higher sensitivity to light elements and magnetic structures, facilitating mode analysis that underpins NTE mechanisms. The volume coefficient β is approximated as β = 3α for cases, derived from parameter variations as β ≈ (Δa/a)/ΔT, with typical resolutions of 10^{-6} K^{-1} or better in laboratory setups. Limitations include the need for crystalline samples and potential at high temperatures, though these techniques excel in confirming NTE . For NTE composites, thermogravimetric analysis (TGA) is often coupled with expansion measurements to evaluate thermal stability alongside dimensional behavior. TGA monitors mass loss under controlled heating, revealing decomposition onset and phase integrity, while integrated thermomechanical analysis (TMA) simultaneously tracks expansion to correlate stability with NTE performance. This combination is crucial for assessing composite durability, as NTE fillers can influence matrix degradation, with resolutions tied to microgram mass sensitivity and sub-micron length detection. Advanced techniques have advanced post-2010 studies of dynamic NTE through time-resolved , capturing transient responses to stimuli like pulses. These methods achieve temporal resolution and angstrom spatial precision, enabling observation of NTE during phase transitions or excitations, such as in nanolayered structures where negative out-of-plane expansion is quantified via transient peak shifts. Challenges persist in anisotropic samples, where preferred orientations complicate data interpretation and require specialized to ensure uniform probing.

Theoretical Models

The formalism provides a foundational thermodynamic framework for understanding , including negative thermal expansion (NTE). The volumetric coefficient \beta is derived from the \beta = \frac{\gamma C_V}{V B_T}, where \gamma is the , C_V is the at constant volume, V is the , and B_T is the isothermal . This expression arises from thermodynamic identities linking the pressure dependence of to volume changes with . Specifically, starting from the \left( \frac{\partial V}{\partial T} \right)_P = -\left( \frac{\partial S}{\partial P} \right)_T and expressing S in terms of phonon frequencies \omega, the \gamma = -\frac{V}{\omega} \frac{d\omega}{dV} quantifies the anharmonic coupling between atomic vibrations and lattice . When \gamma < 0, the overall \beta < 0, leading to NTE, as vibrational frequencies increase with volume expansion, stabilizing a contracted state upon heating. For the linear expansion coefficient \alpha, the simplifies to \alpha = \frac{\beta}{3} = \frac{\gamma C_V}{3 V B_T}. Mode-specific decompositions extend this formalism by attributing contributions to individual phonon modes, where the mode Grüneisen parameter \gamma_q = -\frac{V}{\omega_q} \frac{d\omega_q}{dV} determines the sign of each mode's effect on expansion. In NTE materials, low-frequency modes with \gamma_q < 0 dominate, as their hardening with increasing volume drives contraction. For instance, in ScF₃, first-principles calculations reveal that the two lowest-energy optic modes at approximately 45 and 46 cm⁻¹ exhibit large negative \gamma_q, accounting for the observed NTE in the quasiharmonic approximation. This decomposition highlights how NTE emerges from a weighted sum of mode contributions, \beta = \frac{1}{V B_T} \sum_q \gamma_q C_{V,q}, where C_{V,q} is the mode-specific heat capacity, emphasizing the role of transverse vibrations in open-framework structures. Density functional theory (DFT) enables ab initio predictions of NTE through calculations of phonon dispersions, capturing anharmonic effects via the quasiharmonic approximation. By computing the dynamical matrix at varying volumes and integrating phonon densities of states, DFT determines frequency shifts that yield negative Grüneisen parameters. For example, in ScF₃, DFT phonon dispersions reveal a preponderance of low-energy rigid-unit modes (RUMs) with negative \gamma_q, predicting isotropic NTE coefficients of approximately -10 × 10⁻⁶ K⁻¹ up to 1100 K. This approach has also predicted NTE in hypothetical structures, such as layered perovskites like Ba₃Zr₂S₇, where DFT simulations show tunable negative expansion arising from octahedral tilting modes, with \alpha \approx -5 to -15 × 10⁻⁶ K⁻¹ depending on strain. Such calculations extend to high-throughput screening of semiconductors, identifying candidates with unstable phonons that stabilize into NTE upon thermal perturbation. Molecular dynamics (MD) simulations complement DFT by modeling thermal contraction in flexible frameworks, incorporating explicit beyond the quasiharmonic limit. In materials like ZrW₂O₈, MD trajectories reveal that NTE stems from transverse oscillations of oxygen atoms in WO₄ tetrahedra, leading to a volumetric contraction of approximately -27 × 10⁻⁶ K⁻¹ at 300 K. The quasi-harmonic approximation within MD refines this by iteratively adjusting parameters to minimize , F(V,T) = E(V) + \sum_q \hbar \omega_q(V) \left( \frac{1}{2} + \frac{1}{e^{\hbar \omega_q / kT} - 1} \right), capturing volume-dependent softening that drives contraction in open structures. For ZrV₂O₇, similar simulations quantify three-phonon contributions, showing enhanced NTE from rotational instabilities in pyrochlore frameworks. Recent advances in the 2020s have integrated machine learning (ML) models for high-throughput screening of NTE materials, accelerating discovery beyond traditional DFT or MD. Multi-step ML frameworks, trained on phonon and structural databases, predict NTE in bulk frameworks by correlating negative Grüneisen parameters with geometric descriptors like framework density, identifying candidates such as novel scandium fluorides with \alpha < -20 × 10⁻⁶ K⁻¹. In two-dimensional systems, graph neural networks have discovered anti-Invar materials exhibiting extreme NTE, with \alpha \approx -50 × 10⁻⁶ K⁻¹, by optimizing lattice parameters against thermal strain datasets. A 2025 development includes an AI-based tensor network method from Caltech that accelerates computations of quantum atomic vibrations (phonons) by 1,000 to 10,000 times while maintaining accuracy comparable to quantum MD, aiding predictions of anharmonic effects in NTE materials. However, these models often rely on classical approximations, limiting accuracy in capturing quantum effects like zero-point anharmonicity or tunneling in light-atom frameworks.

Applications

Engineering Uses

Negative thermal expansion (NTE) materials are employed in precision optics for thermal compensation, where dimensional stability is critical over varying temperatures. In , NTE substrates counteract the positive of printed circuit boards (PCBs), reducing stress on components in high-reliability applications. For instance, CERSAT™ ceramic substrates from Nisshinbo provide negative CTE values around -7.0 to -8.2 × 10⁻⁶ K⁻¹, suitable for packaging in where temperature cycles can induce failures. Similarly, ALLVAR Alloy 30, with a CTE of -30 × 10⁻⁶ K⁻¹, is integrated into components like optical benches to maintain alignment under cryogenic conditions. Composite integration leverages NTE materials in bilayer structures with positive-CTE counterparts to achieve zero overall , enhancing structural in demanding environments. These designs, such as those combining NTE fillers like ZrW₂O₈ with metal matrices, are applied in engine components to minimize thermal distortion and in bridge joints to prevent cracking from seasonal temperature swings. In microelectromechanical systems (), NTE metamaterials enable precise actuation and sensing by compensating thermal drifts, with applications in resonant sensors for inertial achieving stability over -40°C to 85°C ranges.

Challenges and Future Directions

Despite their unique properties, negative thermal expansion (NTE) materials face significant material limitations that impede practical implementation. Compounds such as ZrW₂O₈ are notoriously brittle, exhibiting sensitivity to mechanical stress and environmental factors like and , which compromise structural integrity during processing or use. of these materials often requires stringent conditions, including high-purity precursors and controlled s, leading to challenges in and for industrial production. Furthermore, the operational ranges for NTE are typically narrow—for example, ZrW₂O₈ displays NTE from 0.3 to 1050 —limiting applicability in environments with varying thermal demands. Performance gaps persist in achieving consistent and reliable NTE behavior. Isotropic NTE at ambient temperatures is difficult to attain across most materials, though recent advancements in fluorides like MHfF₆ have demonstrated coefficients as low as -7.26 × 10⁻⁶ K⁻¹ from 175 K to 475 K, offering partial solutions. In framework structures, such as metal-organic frameworks, during phase transitions disrupts reversible expansion control, reducing efficiency in dynamic thermal environments. Looking to future directions, research in the 2020s emphasizes innovative material classes and discovery tools to overcome these hurdles. Biomimetic approaches to NTE polymers aim to enhance cryogenic performance and adhesion in composites. AI-driven methods, such as multi-step applied to databases like ICSD, have screened over 1,000 candidates and identified around 57 high-probability NTE materials, establishing scaling relationships for and to guide design. Recent advances as of 2025 include NTE materials for thermal matching in solid oxide fuel cells and cells, improving stability at high temperatures. These trends hold potential for precision thermal management in emerging fields, including , where stable volume control is critical for device integrity. Economic and environmental considerations also shape the trajectory of NTE development. Rare-earth-based materials, such as ScF₃, incur high costs due to scarce elements, constraining scalability and accessibility for commercial applications. To promote , carbon-based alternatives like derivatives and carbon-fiber composites are gaining traction, offering near-zero or negative thermal expansion coefficients (e.g., -8.0 × 10⁻⁶ K⁻¹ at ).

References

  1. [1]
    Negative thermal expansion materials: technological key for control ...
    The mechanisms relevant to NTE are classified into three categories, (i) flexible network, (ii) atomic radius contraction and (iii) magnetovolume effect. An ...
  2. [2]
    Progress of Research in Negative Thermal Expansion Materials
    This report classifies and reviews mechanisms and materials of NTE to suggest means of improving their functionality and of developing new materials.
  3. [3]
    Transverse vibrations driven negative thermal expansion in a ...
    Jul 2, 2008 · Here we proposed an alternative mechanism and show that transverse vibrations at low temperatures, arising from site anisotropy, can induce ...
  4. [4]
    Negative thermal expansion coefficient materials - ScienceDirect.com
    In this review article, we present a general insight into thermal expansion and its physical explanation along with many examples of materials with negative ...
  5. [5]
    Two Decades of Negative Thermal Expansion Research - NIH
    However, the first observation of compounds that contract upon heating dates back several hundred years to the discovery of the “density anomaly of water”.Missing: history | Show results with:history
  6. [6]
    Negative thermal expansion materials - ScienceDirect.com
    The recent discovery of negative thermal expansion over an unprecedented temperature range in ZrW2O8 (which contracts continuously on warming from below 2 K ...
  7. [7]
    Origin of Negative Thermal Expansion in Cubic Revealed by High ...
    May 14, 2001 · Isotropic negative thermal expansion has been reported in cubic Z r W 2 ⁢ O 8 over a wide range of temperatures (0–1050 K).
  8. [8]
    First-Principles Mode Gruneisen Parameters and Negative Thermal ...
    Nov 7, 2012 · Mode Grüneisen parameters were estimated for α − Z r W 2 ⁢ O 8 zone-center modes by means of density functional theory calculations.Abstract · Article Text · Supplemental Material
  9. [9]
    Which phonons contribute most to negative thermal expansion in ...
    Apr 27, 2023 · We show that the important phonons for the negative thermal expansion in this material are those associated with the rigid unit modes (RUMs) and associated ...INTRODUCTION · Analysis of thermal expansion · Analysis of mode Grüneisen...
  10. [10]
    Nuclear quantum effect with pure anharmonicity and the anomalous ...
    Silicon has a peculiar negative thermal expansion at low temperature. This behavior has been understood with a “quasiharmonic” theory.
  11. [11]
    [PDF] Negative thermal expansion
    Jan 14, 2005 · Negative thermal expansion is when a substance contracts on heating, unlike most substances that expand. It is more common in solids, ...<|control11|><|separator|>
  12. [12]
  13. [13]
    Negative thermal expansion in correlated electron system and Fermi ...
    Aug 5, 2025 · From experimental results, we found that many of these materials, such as PrPtBi, Ce0.75La0.25B6, MgB2, Nb show negative thermal expansion (NTE) ...
  14. [14]
  15. [15]
  16. [16]
  17. [17]
    Negative Thermal Expansion from 0.3 to 1050 Kelvin in ZrW2O8
    Negative thermal expansion was found for ZrW2O8 from 0.3 kelvin to its decomposition temperature of about 1050 kelvin. Both neutron and x-ray diffraction ...
  18. [18]
    Negative thermal expansion emerging upon structural phase ...
    ZrV2O7 and HfV2O7, which show negative thermal expansion (NTE) in the high-temperature phase, were investigated using X-ray diffraction and heat capacity ...Missing: discovery date
  19. [19]
    Pronounced Negative Thermal Expansion from a Simple Structure
    The thermal expansion is confirmed to be neg. over most of the temp. range from 5 to 300 K. The main problems with the measurements are the very small ...
  20. [20]
    A Study of KCd[Ag(CN)2]3 | The Journal of Physical Chemistry C
    Mar 4, 2020 · The negative thermal expansion material potassium cadmium dicyanoargentate, KCd[Ag(CN)2]3, is studied at high pressure using a combination ...
  21. [21]
  22. [22]
    Tailoring the thermal and thermomechanical characteristics of novel ...
    The aim of this study is to synthesize novel MAX phase boron (B) composites for high-temperature applications.
  23. [23]
    Thermoelastic study of nanolayered structures using time-resolved X ...
    Jan 14, 2014 · This temperature dependent negative linear thermal expansion in z-direction is consistent with literature values for thin films of PZT and ...Missing: post- | Show results with:post-
  24. [24]
    A high-throughput framework for lattice dynamics - Nature
    Nov 14, 2024 · The volumetric thermal expansion can be directly computed from the total Grüneisen parameter using the following equation: \beta =\frac ...
  25. [25]
    Negative thermal expansion and its relation to high pressures
    Mar 26, 2004 · From equation of state considerations, one can relate NTE to negative Grüneisen parameters (thermal or electronic). Under pressure, these ...Missing: derivation paper
  26. [26]
    Quantitative understanding of negative thermal expansion in ...
    Negative thermal expansion (NTE)—the phenomenon where some materials shrink rather than expand when heated—is both intriguing and useful but remains poorly ...Abstract · Article Text · REAL-SPACE ANALYSIS OF... · LOCAL STRUCTURAL...
  27. [27]
    Tunable Negative Thermal Expansion in Layered Perovskite Ba 3 Zr ...
    May 25, 2025 · The n = 2 Ruddlesden−Popper Ba 3 Zr 2 S 7 sulfide exhibits negative thermal expansion (NTE) that can be tuned by applying pressure.Missing: photoinduced | Show results with:photoinduced
  28. [28]
    High-throughput density-functional perturbation theory phonons for ...
    May 1, 2018 · In this work, we perform ab initio calculations of the full phonon dispersion and vibrational density of states for 1521 semiconductor compounds.Missing: NTE | Show results with:NTE
  29. [29]
    Efficient Calculation of the Negative Thermal Expansion in ZrW2O8
    We present a study of the origin of the negative thermal expansion (NTE) on ZrW2O8 by combining an efficient approach for computing the dynamical matrix ...
  30. [30]
    Anharmonicity in negative thermal expansion materials ZrW2O8 and ...
    Aug 28, 2024 · This work investigates the anharmonic properties of ZrW 2 O 8 and ZrV 2 O 7 , which are typical negative thermal expansion (NTE) materials, based on three- ...
  31. [31]
    Searching for Negative Thermal Expansion Materials with Bulk ...
    Searching for Negative Thermal Expansion Materials with Bulk Framework Structures and their Relevant Scaling Relationships through Multi-step Machine Learning.
  32. [32]
    Machine learning enables the discovery of 2D Invar and anti-Invar ...
    Aug 14, 2024 · While uncommon, some materials expand when cooled, demonstrating negative thermal expansion (NTE). Even more exceptional is zero thermal ...
  33. [33]
    Machine learning unravels quantum atomic vibrations in materials
    Sep 16, 2025 · The new machine learning approach could be extended to compute all quantum interactions, potentially enabling encyclopedic knowledge about how ...
  34. [34]
    Glass ceramic ZERODUR®: Even closer to zero thermal expansion
    Jun 3, 2021 · With LAS glass ceramics such as ZERODUR, the nano-crystals exhibit a negative thermal expansion. A fine balanced mixture of the crystal phase ...
  35. [35]
    Technical properties for ZERODUR® glass-ceramic - SCHOTT
    ZERODUR® glass-ceramic have made it ideal for a vast range of applications that rely on exceptionally low thermal expansion materials.Low-Thermal-Expansion... · Extremely Low Thermal... · Thermal Expansion
  36. [36]
    Negative Thermal Expansion Ceramic Substrate: CERSAT
    Negative Thermal Expansion Ceramic Substrate: CERSAT™ is suitable as a packaging material for parts that need temperature compensation.
  37. [37]
    Thermal Conductivity & CTE of Materials: Can We Engineer Them?
    Aug 30, 2016 · Here is where negative CTE or negative thermal expansion (NTE) materials have been receiving particular attention in the electronics industry ...
  38. [38]
    What is negative thermal expansion? ALLVAR Alloys.
    Negative Thermal Expansion refers to a material that expands when cooled and contracts when heated. ALLVAR Alloys exhibits a -30 ppm/°C CTE.Missing: formula | Show results with:formula
  39. [39]
    [PDF] Engineering Precisely Controlled Negative and Zero Thermal ...
    Negative thermal expansion materials can be incorporated into a PTE matrices to create ZTE or better match the CTE of two materials at an interface. More ...
  40. [40]
    Negative Thermal Expansion Metamaterials: A Review of Design ...
    This review examines a range of fabrication methods, encompassing both additive and traditional techniques, and explores the diverse materials used in the ...
  41. [41]
  42. [42]
  43. [43]
  44. [44]