Fact-checked by Grok 2 weeks ago

Magnetic anisotropy

Magnetic anisotropy is the directional dependence of a material's magnetic properties, arising from the interaction between the and the material's internal structure, such as its crystal lattice or shape, which causes the to prefer alignment along certain "easy" directions over others. This phenomenon is fundamental to ferromagnetic and ferrimagnetic materials, where the energy required to reorient the —known as the magnetic anisotropy energy ()—varies with direction, influencing properties like and . The primary types of magnetic anisotropy include , which originates from spin-orbit coupling between electron spins and the crystal lattice; shape anisotropy, due to the demagnetizing fields from the material's geometry; and stress or magnetoelastic anisotropy, induced by mechanical strain via effects. is an intrinsic property quantified by phenomenological constants (e.g., K_1 and K_2), as seen in where the \langle 111 \rangle direction is easy and \langle 100 \rangle is hard, with K_1 = -1.35 \times 10^5 ergs/cm³ at 300 K. Shape anisotropy dominates in elongated particles like needles, favoring along the long axis to minimize magnetostatic energy, while surface anisotropy becomes significant in thin films and nanostructures. At the atomic level, magnetic anisotropy stems from relativistic spin-orbit interactions, dipole-dipole couplings, and electronic exchange forces, with the MAE often calculated using to predict preferred orientations. In two-dimensional van der Waals materials, it plays a crucial role in stabilizing long-range magnetic order against , as predicted by the Mermin-Wagner theorem, enabling intrinsic 2D through mechanisms like ligand-mediated spin-orbit . Magnetic anisotropy is essential for applications in permanent magnets, where high anisotropy fields (e.g., up to 13 T in YCo₅ alloys) enable strong retention of ; spintronic devices, such as magnetic tunnel junctions relying on perpendicular magnetic anisotropy for ; and emerging quantum technologies in 2D heterostructures, where it can be tuned via gating, strain, or doping for voltage-controlled magnetism. Advances since the have leveraged these effects to develop high-performance magnets, underscoring anisotropy's role in modern materials engineering.

Definition and Fundamentals

Definition

Magnetic anisotropy describes the directional dependence of a material's magnetic properties, including and , on the orientation of the magnetic moments relative to the or the external shape of the sample. This arises because the energy associated with the varies with direction, making certain orientations more favorable than others. In isotropic materials, magnetic properties are uniform in all directions, but anisotropy introduces preferred alignments that influence overall magnetic behavior. In ferromagnetic materials, magnetic anisotropy is crucial for establishing stable directions and enabling loops, as it creates energy barriers that resist changes in direction under applied fields. Without sufficient anisotropy, ferromagnets would exhibit superparamagnetic behavior rather than persistent , lacking the and observed in . This directional preference stabilizes long-range magnetic order by breaking in the system. The landscape of an anisotropic ferromagnet features easy axes, along which the aligns with minimal ; hard axes, requiring significant to orient ; and, in some cases, axes with energies between these extremes. The magnitude of this variation is characterized by the constant K, which quantifies the difference in magnetic per unit volume between easy and hard directions, typically expressed in units of J/m³.

Classification of Types

Magnetic anisotropy is classified primarily according to the of the material's and the resulting energy landscape for , which defines the easy, intermediate, and hard axes along which the prefers to align. This arises from the point groups of the lattice, governed by Neumann's principle, which states that the of any must include the symmetry elements of the . For instance, high-symmetry cubic point groups (e.g., m3m) permit cubic with equivalent directions along body diagonals or face normals, while lower-symmetry hexagonal (e.g., 6/mmm) or orthorhombic (e.g., mmm) point groups lead to uniaxial or biaxial/triaxial forms, respectively, where the energy varies distinctly along one, two, or three principal axes. Uniaxial anisotropy features a single easy , with magnetization energy minimized along that direction and maximized perpendicular to it, resulting in two equivalent energy minima. This type is prevalent in materials with hexagonal crystal symmetry, such as , where the c-axis serves as the easy direction. Magnetization reversal in uniaxial systems typically occurs via coherent rotation of the magnetization , as described by the Stoner-Wohlfarth model, leading to abrupt switching at a critical field angle and characteristic angular dependence of . Biaxial and triaxial anisotropies involve two or three mutually perpendicular principal axes with distinct energy levels, often arising in with orthorhombic or lower , or induced by or interfaces in thin films. Biaxial anisotropy, for example, exhibits two orthogonal axes in the plane, as observed in La0.67Sr0.33MnO3 films on SrTiO3 substrates, where fourfold combines with in-plane to define the axes; reversal proceeds through and domain propagation rather than pure . Triaxial anisotropy, with three unequal axes (, , hard), is exemplified in two-dimensional ferromagnets like CrSBr, where spin-orbit coupling and lattice distortion yield distinct in-plane and out-of-plane preferences, influencing reversal via a combination of and wall motion. Cubic anisotropy reflects the high symmetry of face-centered cubic (FCC) or body-centered cubic (BCC) lattices, featuring multiple equivalent easy directions—six along <100> in BCC iron (due to positive anisotropy constant) or eight along <111> in FCC nickel (negative constant)—with energy barriers separating them. In iron, magnetization reversal involves complex multi-phase processes, classified into four modes (I: linear response; II: partial rotation; III: domain nucleation; IV: saturation), often dominated by 90° or 180° domain wall motion rather than coherent rotation, depending on field orientation relative to the axes.

Physical Origins

Magnetocrystalline Anisotropy

Magnetocrystalline anisotropy arises intrinsically from the interaction between the spin magnetic moments of electrons and the crystal lattice, primarily through spin-orbit coupling and crystal field effects that perturb the atomic orbitals. Spin-orbit coupling links the orbital motion of electrons to their spin, while the crystal field—generated by the surrounding ions—splits the degenerate atomic orbitals, favoring certain directions aligned with the lattice symmetry. This results in a preferred of along specific crystallographic axes, such as the easy axis in hexagonal crystals. Phenomenologically, magnetocrystalline anisotropy is described by the E_a, which depends on the of \mathbf{M} relative to the principal axes. For simple uniaxial , it is often expressed as E_a = K \sin^2 [\theta](/page/Theta), where K is the anisotropy constant and \theta is the angle between \mathbf{M} and the easy axis; more complex cubic symmetries require higher-order terms like K_1 ( \alpha_1^2 \alpha_2^2 + \alpha_2^2 \alpha_3^2 + \alpha_3^2 \alpha_1^2 ) + K_2 \alpha_1^2 \alpha_2^2 \alpha_3^2, with \alpha_i as cosines. These expressions capture how the energy landscape varies with , contributing to the overall magnetic behavior classified under intrinsic types. Early observations of were made through measurements on single crystals, notably by and Kornetzki, who demonstrated direction-dependent in and . Their work highlighted the lattice's role in determining magnetic along specific axes. In key materials like rare-earth compounds, is exceptionally strong due to large orbital moments in rare-earth ions. For instance, Nd₂Fe₁₄B exhibits an anisotropy constant K_1 \approx 4.3 \times 10^7 erg/cm³ at , making it ideal for high-performance permanent magnets. This strength stems from the interplay of Fe-Fe exchange and Nd orbital contributions. The effective decreases with increasing temperature due to thermal agitation, which randomizes spin orientations and reduces the alignment with lattice axes. Below the T_c, the anisotropy constant K(T) follows a power-law decay, often approximated as K(T) \propto M_s^\alpha(T), where M_s(T) is the and \alpha (typically 3 for metals) reflects the of ; at T_c, anisotropy vanishes as ferromagnetic order is lost.

Shape and Magnetoelastic Anisotropy

Shape anisotropy arises from the magnetostatic energy associated with demagnetization fields in non-spherical magnetic particles, where the geometry of the sample influences the preferred direction of . In non-spherical particles, such as ellipsoids, the demagnetization factor N varies along different , leading to lower magnetostatic energy when the magnetization aligns with the direction of the smallest N. For prolate spheroids, which are elongated along one , the demagnetization field is minimized when the magnetization points along the long , thereby favoring this orientation as the easy of . Magnetoelastic anisotropy originates from the coupling between the and through the effect, where mechanical alters the landscape. This extrinsic anisotropy is described by the magnetoelastic term -B \sigma, where B is the magnetoelastic and \sigma is the applied ; the energy varies with the angle between the and stress direction, typically as -B \sigma \cos^2 \theta for uniaxial . In materials exhibiting positive , tensile along the direction lowers the , stabilizing that , while can induce the opposite effect. The effective field H_k, which quantifies the strength required to overcome the anisotropy, is given by H_k = 2K / M_s for uniaxial systems, where K is the constant and M_s is the saturation magnetization. Shape effects modify this through the demagnetization tensor, yielding an effective field H_{k,\text{eff}} = 2K / M_s + (N_\perp - N_\parallel) M_s, where N_\perp and N_\parallel are the demagnetization factors perpendicular and parallel to the easy axis, respectively; this adjustment highlights how can amplify or reduce the total . In thin films and elongated nanostructures, shape anisotropy enhances perpendicular magnetic anisotropy by promoting out-of-plane magnetization alignment due to reduced demagnetization fields in the perpendicular direction. For instance, in NixCo_{100-x} films, elongated nanostructures formed during deposition contribute to strong perpendicular anisotropy, enabling applications in high-density magnetic recording media. In nanoparticles, shape anisotropy often dominates over , particularly in elongated or faceted structures where the magnetostatic contribution exceeds intrinsic crystal effects. For magnetite nanoparticles with moderate elongation, the shape anisotropy energy can be up to five times larger than the magnetocrystalline term, dictating the overall magnetic stability and reversal behavior.

Microscopic Mechanisms

Atomic-Level Contributions

At the atomic level, magnetic anisotropy arises primarily from spin-orbit coupling, a relativistic interaction that couples the spin of an to its orbital . This coupling introduces an energy dependence on the relative orientation of the spin and orbital moments, favoring certain directions due to the atomic environment. In the presence of a field, the degenerate d- or f-orbitals of or rare-earth ions split into levels with different symmetries, leading to preferred spin orientations that minimize the total energy when combined with spin-orbit effects. This splitting breaks the spherical symmetry of the free ion, making the orbital contributions directionally dependent and thus contributing to . Theoretically, this is captured using , where the spin-orbit H_{\mathrm{SO}} = \lambda \mathbf{L} \cdot \mathbf{S} (with \lambda as the , \mathbf{L} the orbital , and \mathbf{S} the spin ) provides a second-order correction to the ground-state . The emerges from the difference in these corrections for spins aligned along different axes, proportional to the expectation value of the orbital moment . In transition metals, involving partially filled d-orbitals, the orbital moments are partially quenched by the crystal field, resulting in relatively weaker anisotropy compared to rare-earth elements with f-electrons. Rare-earth f-electrons, being more localized and less affected by hybridization, retain larger unquenched orbital moments, leading to stronger spin-orbit coupling effects and higher energies. Experimentally, the orbital contributions to atomic-level anisotropy are probed using X-ray magnetic circular dichroism (XMCD), which measures the difference in absorption of circularly polarized X-rays by magnetized samples, allowing separation of spin and orbital moments via sum rules. For instance, XMCD spectra at the L-edges of transition metals reveal how orbital polarization varies with magnetization direction, directly linking to the microscopic .

Molecular-Level Contributions

In coordination compounds, single-ion anisotropy emerges from the coupling of spin-orbit interactions with the crystal field generated by surrounding s, which preferentially stabilizes certain spin orientations in ions, often resulting in easy-axis magnetic behavior. This effect is particularly pronounced in ions with unquenched orbital , such as high-spin Co(II) or Fe(II), where the ligand field splits the d-orbitals, and second-order spin-orbit coupling lifts the degeneracy of the spin multiplet states. Building on atomic-level spin-orbit coupling, the molecular environment further modulates this anisotropy through asymmetric ligand arrangements that enhance . A representative example is the high-spin cluster ([Mn12O12(CH3COO)16(H2O)4]), which possesses a total ground state of S=10 and exhibits strong axial anisotropy dominated by contributions from the eight Mn(III) ions (each S=2). The effective Hamiltonian for this cluster includes the term H = D S_z^2, where D ≈ -0.46 cm⁻¹ quantifies the easy-axis anisotropy along the molecular S4 axis, arising primarily from the Jahn-Teller distortions and ligand field effects on the Mn(III) sites. This axial preference creates a for the , enabling single-molecule magnet behavior at low temperatures. In polynuclear molecular magnets, interactions via bridging ligands propagate single-ion across multiple spin centers, amplifying the overall molecular magnetic hardness. Oxygen- or carboxylate-based bridges, for instance, facilitate antiferromagnetic or ferromagnetic coupling that aligns local anisotropies, as seen in clusters where ligand-mediated integrals depend on orbital overlap and geometry. This transmission can introduce subtle transverse components if the bridges distort the local symmetry, influencing relaxation pathways. The magnitude and nature of molecular anisotropy are precisely characterized by zero-field splitting (ZFS) parameters D and E, incorporated into the general spin Hamiltonian H = D \left(S_z^2 - \frac{S(S+1)}{3}\right) + E(S_x^2 - S_y^2), where D governs the axial splitting (negative for easy-axis systems) and E the rhombic distortion. These parameters originate from second-order perturbations involving spin-orbit coupling and ligand field stabilization energies, with typical values for transition metal complexes ranging from |D| ~ 1–100 cm⁻¹ depending on the ion and coordination. In practice, ab initio calculations and electron paramagnetic resonance spectroscopy are used to extract D and E, revealing how ligand choice tunes the anisotropy tensor. For molecular nanomagnets, the ZFS-induced anisotropy creates an energy barrier to magnetization reversal of Δ = |D| S², which hinders thermal relaxation and suppresses quantum tunneling between opposite spin projections at cryogenic temperatures. This barrier, for instance, reaches ~25 K in Mn12-acetate (with S=10), enabling hysteresis and potential applications in data storage, though transverse terms from E or higher-order anisotropy can reduce the effective blocking by promoting tunneling. Optimizing ligand fields to maximize |D| while minimizing E remains key to enhancing these barriers in designer systems.

Anisotropy in Single-Domain Systems

Uniaxial Anisotropy

Uniaxial represents the simplest form of magnetic in single-domain ferromagnetic particles, characterized by a single preferred direction, or easy , along which the aligns to minimize , with the perpendicular to this being degenerate and harder. This configuration arises in elongated particles or crystals with suitable symmetry, leading to distinct reversal behaviors under applied fields. The total associated with uniaxial is given by E = K V \sin^2 \theta, where K is the , V is the particle volume, and \theta is the angle between the vector and the easy ; positive K favors alignment along the easy . The Stoner-Wohlfarth model provides a foundational description of magnetization switching in such systems, assuming coherent rotation of the uniform without domain wall nucleation. Developed in 1948, the model predicts switching under applied magnetic fields via an astroid-shaped critical curve in the plane of applied field components parallel and perpendicular to the easy axis, delineating stable magnetization states. For a field applied along the easy axis, the switching field is H_{sw} = 2K / M_s, where M_s is the saturation , marking the coercivity in ideal cases. Representative examples of uniaxial anisotropy include arrays of nanowires, where shape anisotropy from high aspect ratios dominates, yielding effective uniaxial behavior with coercivities up to several depending on diameter and length. Similarly, (\alpha-Fe_2O_3) particles exhibit uniaxial anisotropy due to their crystallographic structure, influencing and in natural assemblages. In perfect uniaxial systems, hysteresis loops are rectangular, with equal to saturation magnetization and sharp switching at the , reflecting the binary easy-axis preference.

Triaxial Anisotropy

Triaxial anisotropy in single-domain magnetic systems features three mutually perpendicular principal axes characterized by distinct energy preferences: an easy axis for minimum energy, an intermediate axis with moderate energy, and a hard axis requiring maximum energy for magnetization alignment. This low-symmetry configuration arises primarily from orthorhombic crystal structures or engineered geometries that break the rotational invariance present in higher-symmetry cases. The anisotropy energy density is commonly expressed in spherical coordinates (\theta, \phi), where \theta is the polar angle from the easy axis and \phi is the azimuthal angle, as E = K_1 \sin^2 \theta + K_2 \sin^4 \theta \sin^2 2\phi, with K_1 > 0 favoring the easy axis and K_2 > 0 introducing in-plane distinction between intermediate and hard directions; alternatively, it takes a general quadratic form E = K_x m_x^2 + K_y m_y^2 + K_z m_z^2 involving three independent constants K_x, K_y, K_z (where m_i are direction cosines) to describe the ellipsoidal energy surface. The equilibrium paths in triaxial systems are influenced by saddle points in the energy landscape, particularly along the intermediate axis, which exhibits under perturbations or applied fields. These saddle points create multiple low-energy trajectories for , contrasting with the single symmetric path in uniaxial systems. When subjected to external fields, the intermediate axis acts as a metastable , facilitating transitions via or coherent over the saddle, leading to asymmetric switching behavior dependent on field orientation. Magnetization reversal in triaxial single-domain particles often proceeds through non-uniform modes such as , where the rotates in a vortex-like to reduce energy, or , involving localized bending to accommodate the triaxial constraints and lower the reversal barrier. These mechanisms are prominent in elongated or plate-like particles with orthorhombic , where the competition between , demagnetization, and fields dictates the dominant mode; for instance, dominates in prolate shapes under fields near the hard axis. Representative materials exhibiting triaxial anisotropy include orthorhombic crystals like rare-earth orthoferrites (e.g., GdFeO_3), where crystal-field effects on Fe^{3+} ions induce the three distinct axes, resulting in weak with reorientation transitions. Engineered magnetic multilayers, such as strain-tuned Co/Pt heterostructures, can also be designed to mimic triaxial through interfacial effects and substrate-induced distortions, enabling tunable switching for spintronic devices. Compared to uniaxial anisotropy, triaxial systems offer broader energy basins with additional shallow minima near the intermediate axis, promoting more complex and potentially multistep reversal processes that enhance stability against while complicating field-driven control. Triaxial anisotropy extends uniaxial by incorporating a secondary transverse term that differentiates azimuthal directions.

Cubic Anisotropy

Cubic anisotropy arises in single-domain ferromagnetic materials with cubic symmetry, such as body-centered cubic (bcc) iron and face-centered cubic (fcc) , where the magnetocrystalline depends on the orientation of the relative to the high-symmetry axes. This form of anisotropy is characterized by equivalent <100>, <110>, and <111> directions, leading to multiple easy, intermediate, and hard axes. The landscape favors certain high-symmetry directions due to spin-orbit , influencing the material's response to external fields and its suitability for applications in and sensors. The leading terms of the magnetocrystalline anisotropy energy density for cubic crystals are given by \frac{E}{V} = K_1 (\alpha_1^2 \alpha_2^2 + \alpha_2^2 \alpha_3^2 + \alpha_3^2 \alpha_1^2) + K_2 \alpha_1^2 \alpha_2^2 \alpha_3^2, where \alpha_1, \alpha_2, \alpha_3 are the direction cosines of the vector with respect to the cubic axes, and K_1 and K_2 are the first- and second-order anisotropy constants, respectively. This expression captures the fourth-order of the cubic , with the K_1 term dominating the overall behavior. The easy directions depend primarily on the sign of K_1: for K_1 > 0, the <100> directions are easiest, while for K_1 < 0, the <111> directions are favored. In bcc iron, K_1 is positive (approximately $4.8 \times 10^4 J/m³ at ), resulting in easy axes along <100>. Conversely, in fcc , K_1 is negative (approximately -0.5 \times 10^4 J/m³), making <111> the easy directions. The higher-order constant K_2 plays a crucial role in resolving degeneracies among the easy axes when the K_1 term alone leaves multiple directions equivalent, such as lifting the sixfold degeneracy of <100> or <111> axes by introducing finer energy distinctions. For most cubic ferromagnets like iron and , |K_2| is much smaller than |K_1| (e.g., K_2 \approx 0.15 \times 10^4 J/m³ in ), so it provides only subtle corrections without altering the primary easy directions. However, in materials where K_2 is comparable or larger, it can stabilize intermediate directions like <110> or induce more complex anisotropy landscapes. Under applied , cubic anisotropy leads to field-induced transitions where the rotates coherently from easy toward hard directions, particularly when the field is aligned with a hard . This rotation occurs to minimize the total , which includes both the anisotropy and Zeeman terms, resulting in nonlinear curves and responses. For instance, in iron, a field along a <111> hard causes progressive rotation of the from <100> toward the field direction, enabling only at higher fields compared to easy-axis alignment. Such behavior is central to understanding and switching in cubic single-domain particles.

Advanced Topics and Applications

Anisotropy in Low-Dimensional Materials

In low-dimensional materials, magnetic manifests distinctly due to reduced symmetry, enhanced surface effects, and quantum confinement, leading to novel phenomena not observed in systems. Perpendicular magnetic (PMA) in ultrathin films, typically below 2 nm thickness, arises primarily from interface effects at the ferromagnet/oxide boundary, where orbital hybridization and spin-orbit coupling (SOC) contribute to a large effective anisotropy constant K_{\text{eff}}. For instance, in Co/Pt multilayers or CoFeB/MgO structures, these interfacial contributions can yield K_{\text{eff}} > 10^6 erg/cm³, surpassing the shape limit and stabilizing out-of-plane even in films where demagnetization would otherwise favor in-plane . This enhancement is for applications in high-density , as it allows single-domain-like behavior in nanoscale volumes. Two-dimensional van der Waals (vdW) magnets exemplify strong out-of-plane anisotropy driven by both interlayer magnetic coupling and intrinsic SOC. In materials like CrI₃, the monolayer exhibits Ising-type ferromagnetism with perpendicular spin orientation, stabilized by SOC at the heavy iodine sites, while bilayer stacking introduces antiferromagnetic interlayer coupling that can be tuned to ferromagnetic via electrostatic gating or strain, yielding anisotropy energies on the order of several meV per spin. Similarly, Fe₃GeTe₂, a metallic vdW ferromagnet, displays robust PMA with Curie temperatures above 200 K, where SOC proximity from adjacent layers or substrates further amplifies the out-of-plane preference, enabling tunable domain walls for spintronic devices. These systems highlight how weak vdW interlayer interactions, combined with SOC, preserve long-range order down to the atomic limit. Nanoparticle assemblies introduce superstructure , where collective arrangements amplify effective magnetic fields beyond individual particle contributions. A 2025 study on chains of Fe₃O₄ demonstrated that linear superstructures induce a uniaxial anisotropy field exceeding 10 kOe, attributed to dipolar interactions and shape-induced alignment during , resulting in enhanced and suitable for permanent magnet nanocomposites. This superstructure effect scales with chain length and packing density, offering a pathway to engineer macroscopic anisotropy from mesoscale building blocks. Quantum confinement in one-dimensional (1D) nanowires further intensifies anisotropy through shape dominance and electronic subband formation. In ferromagnetic nanowires like α-Fe₂O₃ or FeGe₂, the high promotes strong shape anisotropy along the wire axis, with effective fields up to several kOe, while quantum confinement lifts orbital degeneracies, enhancing SOC-driven contributions. This builds briefly on classical shape anisotropy principles but is amplified by 1D band structure effects, enabling applications in high-sensitivity magnetic sensors. Recent developments have focused on probing subtle in-plane in systems, revealing ultraweak fields that challenge detection limits. A study in PNAS used tunneling magneto-conductance measurements to detect ultraweak in-plane fields as low as ~1 mT in micrometer-scale antiferromagnets such as CrCl₃, uncovering hidden easy axes due to distortions and enabling precise control for processing. These advances underscore the potential of low-dimensional platforms for discovering emergent magnetic phases.

Measurement Techniques and Spintronic Applications

Torque magnetometry measures the torque exerted on a magnetized sample in an applied magnetic field, providing a direct method to determine the magnetic anisotropy constant K by analyzing the angular dependence of the torque. This technique is particularly effective for single crystals and thin films, where the torque \tau = -\frac{\partial E}{\partial \theta} relates to the anisotropy energy E, allowing extraction of uniaxial or higher-order constants through sinusoidal fitting of torque curves at various field strengths. Ferromagnetic (FMR) probes the of under , enabling determination of effective anisotropy fields H_k from shifts in frequency. In perpendicular geometry, the Kittel equation \omega = \gamma \sqrt{(H + H_k)(H + H_k + 4\pi M_s)} (where \gamma is the and M_s the ) is fitted to frequency-field data, distinguishing contributions from , magnetocrystalline, and . FMR extends this to thin films and multilayers, revealing and gradients with sensitivities down to mT fields. Vibrating sample magnetometry (VSM) quantifies through hysteresis loops, measuring as a function of applied field to observe switching behavior and . By rotating the sample and recording loop shapes, effective is inferred from the angular variation in and , often using Stoner-Wohlfarth models for single-domain particles. VSM's high sensitivity (down to $10^{-6} emu) suits room-temperature studies of ferromagnetic materials, complementing torque methods for polycrystalline samples. Angular-dependent hysteresis loops, typically acquired via VSM or magneto-optic , allow extraction of anisotropy fields H_k by fitting loop parameters to models like the coherent model. For uniaxial , the hard-axis loop width approximates H_k, while full angular scans (0° to 180°) enable least-squares fitting to \sin^2\theta dependencies, yielding H_k = 2K / M_s. This approach is robust for thin films, resolving H_k values from 10 to several kOe with errors below 5%. In , perpendicular magnetic anisotropy (PMA) in magnetic tunnel junctions (MTJs) enhances thermal stability and enables high-density magnetoresistive (MRAM) by favoring out-of-plane magnetization over in-plane, reducing the need for complex toggle switching mechanisms in early in-plane designs. PMA-engineered CoFeB/MgO interfaces achieve anisotropy fields exceeding 5 kOe. Anisotropy tuning stabilizes —topological spin textures—for , where interfacial PMA confines skyrmions to nanoscale tracks, preventing annihilation at edges. By introducing Dzyaloshinskii-Moriya interaction gradients via material stacking, skyrmion diameters are reduced to 5-10 , enabling current-driven motion at velocities up to 100 m/s with pinning energies below 1 eV, critical for multi-bit storage densities exceeding 1 Tb/in². High-anisotropy barriers at track edges further suppress skyrmion collapse under . Recent advances include field-free switching in PMA systems using anisotropy gradients induced by oblique deposition, where tilted evaporation creates lateral variations in interfacial anisotropy, generating an effective switching field of ~50 without external bias. A 2025 study on obliquely deposited Ta/CoFeB/Pt and W/CoFeB/Pt heterostructures demonstrated deterministic toggling at , leveraging type X/Y current-induced switching via coherent rotation and domain-wall propagation for energy-efficient spintronic logic.

References

  1. [1]
    3. Magnetic Anisotropy | College of Science and Engineering
    Magnetocrystalline anisotropy is the energy necessary to deflect the magnetic moment in a single crystal from the easy to the hard direction.Missing: fundamentals | Show results with:fundamentals
  2. [2]
    Magnetic Anisotropy - an overview | ScienceDirect Topics
    Magnetic anisotropy is defined as the exhibition of differing magnetic properties along one or more axes in a ferromagnetic material.Publisher Summary · Monolayer Films: Magnetism · 4.1. 2 Early Work<|control11|><|separator|>
  3. [3]
    Magnetic Anisotropy in 2D van der Waals Magnetic Materials and ...
    Aug 5, 2025 · Magnetic anisotropy is a fundamental and essential physical property of magnetic materials, describing the tendency of magnetizations to ...
  4. [4]
    Magnetic Anisotropy - an overview | ScienceDirect Topics
    Magnetic anisotropy is the directional dependence of a material's magnetic moment, where the energy required to deflect it from its preferred direction.
  5. [5]
    None
    Summary of each segment:
  6. [6]
    [PDF] Minimum Anisotropy of a Magnetic Nanoparticle out of Equilibrium
    The abrupt change in the hysteresis abundance among these metals suggests a concept of minimum magnetic anisotropy required for magnetic hysteresis, which is ...<|control11|><|separator|>
  7. [7]
    Determination of anisotropy constants via fitting of magnetic ...
    Aug 5, 2021 · The Stoner–Wohlfarth model can be used to numerically fit measured magnetic hysteresis curves to more easily determine these anisotropy constants.
  8. [8]
    [PDF] 3 Magnetic Anisotropy
    Jul 3, 2019 · Magnetic anisotropy - symmetry of crystals. ○. Neumann's Principle: The symmetry elements of any physical property of crystal must include ...
  9. [9]
  10. [10]
    Room temperature biaxial magnetic anisotropy in La 0.67 Sr 0.33 ...
    Aug 3, 2018 · The angular dependent magnetic anisotropy symmetry in our STO buffered LSMO films shows dominant biaxial anisotropy even at RT, enhanced by ...
  11. [11]
    Triaxial magnetic anisotropy in the two-dimensional ferromagnetic ...
    Oct 21, 2021 · Our results show that bulk CrSBr is a magnetic semiconductor and has the easy magnetization b axis, hard c axis, and intermediate a axis. Thus ...Missing: examples | Show results with:examples
  12. [12]
  13. [13]
    Anisotropy - DoITPoMS
    Shape Anisotropy in Polycrystals. If a sample is non-spherical it is easier to magnetise with the dipole moments pointing along a long axis. The field required ...
  14. [14]
    Anisotropy in magnetoelastic effects of | Phys. Rev. Materials
    Nov 4, 2024 · We develop the phenomenological theory for uniaxial ferromagnets and derive explicit formulas for their magnetostriction (MS) and magnetic anisotropy energy ( ...Missing: sigma | Show results with:sigma
  15. [15]
    Influence of stresses and magnetostriction on the soft magnetic ...
    As is known, the stress-induced magnetoelastic anisotropy is expressed by K = −3σλS/2, where σ and λS represent stress and saturation magnetostriction constant, ...
  16. [16]
    Influence of Shape Anisotropy on Magnetization Dynamics Driven ...
    Dec 8, 2016 · In this paper, we introduce the analytic demagnetizing factors for elliptic film nanomagnets and theoretically investigate the geometric ...Missing: H_k = 2K/ M_s
  17. [17]
    NiCo films with perpendicular magnetization anisotropy deposited ...
    ▻ Meanwhile, the elongated nanostructures are formed in these films, some of them are assembled along the direction pointing out of the substrate, which may ...
  18. [18]
    The role of faceting and elongation on the magnetic anisotropy of ...
    Feb 17, 2020 · ... magnetocrystalline anisotropy and so shape effects are expected to dominate the overall magnetic anisotropy of magnetite nanoparticles. In ...
  19. [19]
    None
    Nothing is retrieved...<|control11|><|separator|>
  20. [20]
    Rare-earth magnetic anisotropy: Is the crystal field theory valid?
    It was suggested in recent papers that the crystal field parameters describing the rare-earth magnetic anisotropy in metals should strongly depend on the ...Missing: splitting preferred
  21. [21]
    [PDF] Crystal fields. Spin-Orbit Coupling Magnetic Anisotropy.
    Sep 5, 2023 · Zeeman interaction. magnetic properties are solved by perturbation methods. The spin moment is coupled to the electron cloud (orbitals): S-O ...
  22. [22]
    Calculations of anisotropic magnetic properties using spin-orbit ...
    Nov 19, 2020 · We analyze several methods of obtaining the accurate relativistic total energy (TE) variations using traditional perturbation theories (PTs) ...Article Text · INTRODUCTION · METHODS · RESULTS AND DISCUSSION
  23. [23]
    Electronic Structure and Magnetic Anisotropy of Single-Layer Rare ...
    Jun 4, 2020 · Besides the TMC, rare-earth compounds are also of particular importance in modern magnets. Rare-earth magnets are well known to be strong ...
  24. [24]
    Rare earth elements and magnetism in metallic systems
    The magnetic moment is usually maintained in a solid where a given rare earth element atom interacts and shares electrons with neighboring atoms.
  25. [25]
    Pinned orbital moments – A new contribution to magnetic anisotropy
    May 6, 2016 · The XMCD effect is related to a change in the scalar product between the atomic magnetic moments and the axial vector of the circular light ...
  26. [26]
    [PDF] Single-Ion Magnetic Anisotropy and Isotropic Magnetic Couplings in ...
    Jul 30, 2013 · The single-ion anisotropy arises from the joint effects of the ligand field and spin−orbit coupling (SOC); these effects are mainly determined ...Missing: seminal | Show results with:seminal
  27. [27]
    Single-Ion Magnets with Giant Magnetic Anisotropy and Zero-Field ...
    Nov 15, 2022 · As a result, some molecular nanomagnets, expected to exhibit large anisotropy energy due to precalculated large splitting parameters, may in ...Introduction · In the Pursuit of the Largest... · Critical Theoretical Aspects
  28. [28]
    spin Hamiltonian for magnetic clusters - American Institute of Physics
    magnetic clusters are Mn12-acetate, whose complete chemi- cal formula is. Mn12O12(CH3COO兲16共H2O兲4, and. Fe8O2共OH兲12共tacn兲6Br8•9H2O. 共where tacn⫽1,4,7 ...
  29. [29]
    New mechanism of kinetic exchange interaction induced by strong ...
    Apr 21, 2016 · Here we show that the exchange interaction between strongly anisotropic doublets of lanthanides, actinides and transition metal ions with unquenched orbital ...
  30. [30]
    Probing Superexchange Interaction in Molecular Magnets by Spin ...
    Superexchange, a long-range interaction, was studied in CoPc films using spin-flip spectroscopy and microscopy, identifying mediating molecular orbitals.
  31. [31]
    Rapid and precise determination of zero-field splittings by terahertz ...
    Apr 19, 2017 · ZFS originates from spin–spin interactions mediated by the ligand field and from spin–orbit coupling. The behavior of ZFS can be adequately ...Epr Spectra In Zero And... · High-Spin Fe(iii): Spin-5/2... · High-Spin Ni(ii): Spin-1...
  32. [32]
    [PDF] Zero-field splitting in transition metal complexes: Ab initio ...
    Dec 2, 2021 · Zero-field splitting (ZFS) is a concept that is commonly used to indicate the loss of the degeneracy of the spin components of a spin-orbit ...<|separator|>
  33. [33]
    A Criterion for the Anisotropy Barrier in Single-Molecule Magnets
    A criterion for the height of the anisotropy barrier Δ = |D|S 2 in single-molecule magnets, and therefore the blocking temperature TB ∝ Δ, is presented.
  34. [34]
    Single-Ion Magnets with Giant Magnetic Anisotropy and Zero-Field ...
    The present review presents the comprehensive state of the art in the topic of single-ion magnets possessing an eminent magnetization-reversal barrier.
  35. [35]
    Fundamental inequalities in the Stoner-Wohlfarth model | Phys. Rev. B
    (1). while the energy associated with the uniaxial magnetic anisotropy of a particle is. E a = k eff V p sin 2 ( θ ) ,. (2). where μ 0 is the magnetic ...
  36. [36]
    A mechanism of magnetic hysteresis in heterogeneous alloys
    A mechanism of magnetic hysteresis in heterogeneous alloys. Edmund Clifton Stoner.
  37. [37]
    High Energy Product Developed from Cobalt Nanowires - Nature
    Jun 18, 2014 · Cobalt nanowires with high aspect ratio have been synthesized via a solvothermal chemical process. Based on the shape anisotropy and ...
  38. [38]
    Magnetic Domain State and Anisotropy in Hematite (α‐Fe2O3) From ...
    Nov 25, 2021 · Stoner and Wohlfarth (1948) theory is used widely to understand the hysteresis behavior of randomly oriented uniaxial SD particle assemblages.
  39. [39]
  40. [40]
    Cubic Anisotropy - an overview | ScienceDirect Topics
    (i) Uniaxial anisotropy occurs in hexagonal crystals such as cobalt. (2) E ... In Fe, magnetocrystalline anisotropy has cubic symmetry, with positive K1.
  41. [41]
    [PDF] Where is magnetic anisotropy field pointing to? - arXiv
    Dec 26, 2013 · The desired result of magnetic anisotropy investigations is the determination of value(s) of various anisotropy constant(s). This is.Missing: triaxial sin²θ + sin⁴θ sin²2φ
  42. [42]
    (PDF) Magnetocrystalline anisotropy energy in cubic Fe, Co, and Ni
    Aug 7, 2025 · PDF | We report an ab initio investigation of the magnetocrystalline anisotropy energy (MAE) in bcc Fe and fcc Co and Ni.
  43. [43]
    Anomalous enhancement in interfacial perpendicular magnetic ...
    Jun 17, 2014 · ... anisotropy energy per unit volume, Keff, for the film was estimated around 5.5 × 106 erg/cm3. Surprisingly, post-annealing at 350°C yielded ...Missing: cm³ | Show results with:cm³
  44. [44]
    Ultrathin perpendicular magnetic anisotropy CoFeB free layers for ...
    Dec 19, 2019 · An ultrathin CoFeB layer with large anisotropy, high saturation magnetization, and good annealing stability to temperatures upwards of 400 °C.Missing: elongated | Show results with:elongated<|separator|>
  45. [45]
    [PDF] Layer-dependent Ferromagnetism in a van der Waals Crystal down ...
    Monolayer chromium triiodide (CrI3) is an Ising ferromagnet with out-of-plane spin orientation, showing ferromagnetic ordering down to the monolayer limit.
  46. [46]
    Drastic change of magnetic anisotropy in Fe 3 GeTe 2 and ... - Nature
    Sep 2, 2021 · We show that perpendicular electric field dramatically enhances the in-plane and out-of-plane magnetic anisotropies in Fe 3 GeTe 2 and Fe 4 GeTe 2 monolayers.
  47. [47]
    Spin–orbit coupling proximity effect in MoS2/Fe3GeTe2 ...
    Jan 24, 2022 · With its strong spin–orbit coupling (SOC), MoS 2 effectively alters the crystal field of the adjacent FGT and its magnetic anisotropy.
  48. [48]
    Superstructure magnetic anisotropy in Fe 3 O 4 nanoparticle chains
    Jul 1, 2025 · This study demonstrates the feasibility and approaches to create “patterned” high magnetic anisotropy in nanoparticle superstructures/assemblies.
  49. [49]
    Magnetic anisotropy of vertically aligned -Fe2O3 nanowire array
    These distinctive magnetic properties are likely related to the surface spins and/or shape effect of nanowires. ... crystalline anisotropy dominates the shape ...
  50. [50]
    Dimensionality Effects in FeGe2 Nanowires: Enhanced Anisotropic ...
    Aug 2, 2017 · Magnetic measurements of FeGe2 nanowire arrays show strongly anisotropic behavior. Temperature and field dependent magnetization and ac ...<|control11|><|separator|>
  51. [51]
    Probing ultraweak in-plane magnetic anisotropy within a two ... - PNAS
    To summarize, probing magnetic anisotropy, particularly ultraweak in-plane magnetic anisotropy, in micrometer-sized 2D magnetic insulators can be extremely ...
  52. [52]
    TORQUE MAGNETOMETRY
    Torque magnetometry is the most straightforward technique to derive the magnetic anisotropy energy of a magnetized body and determine its directions.
  53. [53]
    Resonant torsion magnetometry in anisotropic quantum materials
    Sep 28, 2018 · An established and highly sensitive technique to directly probe small anisotropies in correlated metals and exotic magnets is torque magnetometry.
  54. [54]
    Ferromagnetic resonance spectroscopy for assessment of magnetic ...
    Dec 28, 2006 · Ferromagnetic resonance spectroscopy (FMR) can be used to measure the effective magnetic field within a sample, including the contributions ...
  55. [55]
    Ferromagnetic resonance and control of magnetic anisotropy by ...
    Jan 25, 2019 · FMR is not only one of the most efficient and powerful techniques to observe the MA [32] , but is also used for spin pumping to inject spin ...
  56. [56]
    VSM - MSU Magnetic Nanostructures | Montana State University
    Vibrating sample magnetometry (VSM) is a method by which hysteresis loops (m vs. H) may be accurately measured. The process is to apply a field across a ...
  57. [57]
  58. [58]
    Temperature dependence of magnetization reversal and angular ...
    Sep 27, 2006 · The data needed to fit the model were accumulated by both traditional hysteresis loops taken with a superconducting quantum interference device ...
  59. [59]
    [PDF] Toggle Spin-Orbit Torque MRAM with Perpendicular Magnetic ...
    In particular, STT-MRAM with perpendicular magnetic anisotropy (PMA) is preferred over in-plane anisotropy due to its higher density5 and increased thermal ...
  60. [60]
    Interfacial Perpendicular Magnetic Anisotropy in Sub-20 nm Tunnel ...
    Magnetic tunnel junctions (MTJs) with interfacial perpendicular magnetic anisotropy (PMA) attract much attention due to their utilization in spin-transfer ...
  61. [61]
    An Improved Racetrack Structure for Transporting a Skyrmion - Nature
    Mar 30, 2017 · Here we propose a new skyrmion-based racetrack structure by adding high-K materials (materials with high magnetic crystalline anisotropy) at the edges.Missing: stabilization | Show results with:stabilization
  62. [62]
    Controlled anisotropic dynamics of tightly bound skyrmions in a ...
    Mar 5, 2019 · In this work, we study the influence of induced in-plane anisotropy on a skyrmion in a defect-free environment, at driving current densities of ...
  63. [63]
    Uniaxial Magnetic Anisotropy and Type-X/Y Current-Induced ... - arXiv
    Oct 17, 2025 · Our results show that combining oblique-angle deposition with easy-axis engineering enables deterministic, field-free switching, paving the way ...Missing: gradients | Show results with:gradients