Fact-checked by Grok 2 weeks ago

Magnetization

Magnetization is a quantity that represents the per unit volume in a , quantifying the of aligned magnetic dipoles within it. It arises primarily from the orbital motion and intrinsic of electrons in atoms, where these microscopic currents generate magnetic moments that can align under an applied . In materials, magnetization manifests differently based on the type: in diamagnetic substances, it opposes the field weakly; in paramagnetic ones, it aligns partially with the field; and in ferromagnetic materials, strong internal interactions lead to spontaneous alignment and remanent magnetization even without an external field. The magnetization \mathbf{M} contributes to the total magnetic field \mathbf{B} inside a through the \mathbf{B} = \mu_0 (\mathbf{H} + \mathbf{M}), where \mathbf{H} is the magnetic field strength due to free currents and \mu_0 is the permeability of free space. This formulation separates the effects of external sources (\mathbf{H}) from the 's response (\mathbf{M}), with \mathbf{M} analogous to electric polarization \mathbf{P} in . Magnetization can be induced by external fields or occur spontaneously in certain materials below their , influencing properties like permeability and . Key aspects of magnetization include its measurement in amperes per meter (A/m) and its role in technologies such as , transformers, and MRI imaging, where controlled alignment enhances performance. In ferromagnetic materials, domains of aligned moments form to minimize energy, and loops describe the reversible and irreversible changes during field cycling. Understanding magnetization is fundamental to , bridging microscopic quantum effects with macroscopic applications.

Fundamentals

Definition

Magnetization, denoted as \mathbf{M}, is a vector field that quantifies the magnetic dipole moment density within a material, representing the net alignment of microscopic magnetic moments on a macroscopic scale. It is defined as the total magnetic dipole moment \mathbf{m} per unit volume V, expressed mathematically as \mathbf{M} = \frac{\mathbf{m}}{V}, where \mathbf{M} points in the direction of the net magnetic moment. In the International System of Units (SI), magnetization has units of amperes per meter (A/m), reflecting its nature as a current density equivalent arising from aligned atomic dipoles. The magnitude and direction of \mathbf{M} arise from the collective behavior of atomic or molecular magnetic moments, such as those from electron spins or orbital motions, averaged over the material's volume. This vectorial quantity captures how materials respond to external magnetic influences by developing an internal magnetization that can enhance, oppose, or sustain magnetic fields. Magnetization \mathbf{M} is distinct from the magnetic field strength \mathbf{H}, which originates primarily from free currents and external sources, and the magnetic flux density \mathbf{B}, which includes both vacuum and material contributions to the overall magnetic field; specifically, \mathbf{M} embodies the material's intrinsic response to magnetization processes. The concept of magnetization was formalized by James Clerk Maxwell in the 19th century within his framework of continuum , as detailed in his seminal work A Treatise on Electricity and Magnetism (1873), where it served as a key descriptor for magnetic phenomena in extended media.

Units and Measurement

Magnetization is quantified in the (SI) as amperes per meter (A/m), representing per unit volume. In the older centimeter-gram-second (cgs) electromagnetic unit (emu) system, it is expressed in electromagnetic units per cubic centimeter (emu/cm³), a convention that persists in some historical and specialized literature for comparisons with legacy data. The conversion between these systems follows from the relation involving the permeability of free space, μ₀ = 4π × 10⁻⁷ henries per meter (H/m), where 1 emu/cm³ corresponds to 10³ A/m for magnetization. Experimental measurement of magnetization relies on techniques that detect the magnetic moment induced or present in a sample under controlled fields. Vibrating sample magnetometry (VSM), developed by Simon Foner in 1959, vibrates the sample in a magnetic field to induce a voltage in pickup coils via Faraday's law, enabling precise determination of magnetization versus applied (M vs. H) curves for a wide range of materials. Superconducting quantum interference device () magnetometry offers exceptional sensitivity for low-field measurements, detecting magnetic moments as small as 10⁻⁸ by exploiting quantum interference in superconducting loops, making it ideal for weakly magnetic or nanoscale samples. Torque magnetometry, suitable for anisotropic materials, measures the torque exerted on a sample in a known to infer magnetization direction and magnitude, particularly useful for single crystals where easy and hard axes differ significantly. Calibration of these instruments typically uses standard samples like high-purity cylinders, whose magnetization is known from independent or theoretical calculations, ensuring traceability to units. Error sources in measurements include sample effects, such as demagnetization fields in non-ellipsoidal shapes, which can distort the internal field and lead to inaccuracies in inferred magnetization by up to 10-20% for thin disks or rods unless corrected via shape anisotropy factors. Additional uncertainties arise from positioning errors in the field uniform region or thermal drifts, mitigated by automated alignment and temperature stabilization in modern systems.

Microscopic Origins

Atomic Magnetic Moments

Atomic magnetic moments arise primarily from the orbital and angular momenta of electrons in atoms, serving as the microscopic building blocks of magnetization in materials. The orbital contribution originates from the motion of an electron around the , generating a that produces a magnetic dipole moment. This moment is given by \mu_L = -\frac{[e](/page/Elementary_charge)}{2m_e} \mathbf{L}, where e is the elementary charge, m_e is the , and \mathbf{L} is the orbital . In , \mathbf{L} is quantized, with its magnitude \sqrt{l(l+1)}\hbar for orbital l, leading to discrete possible values for the orbital moment. The angular momentum of an provides another source of , distinct from its orbital motion. The magnetic moment is \mu_S = -g_s \mu_B \frac{\mathbf{S}}{\hbar}, where \mathbf{S} is the angular momentum, \mu_B = \frac{e\hbar}{2m_e} is the (approximately $9.274 \times 10^{-24} J/T), and g_s \approx 2.0023 is the electron g-factor, reflecting the relativistic nature of the electron's intrinsic . For a single , the moment magnitude is about $1.73 \mu_B when considering the effective value g_s \sqrt{s(s+1)} \mu_B with s = 1/2. In multi-electron atoms, the total angular momentum \mathbf{J} = \mathbf{L} + \mathbf{S} determines the net magnetic moment, \mu_J = -g_J \mu_B \frac{\mathbf{J}}{\hbar}, where g_J is the given by g_J = 1 + \frac{J(J+1) + S(S+1) - L(L+1)}{2J(J+1)}. This formula, derived from the vector model of the atom, accounts for the coupling between orbital and contributions, with g_J ranging from 1 (pure orbital) to 2 (pure ). However, in solid-state environments, fields often quench the orbital , reducing \mathbf{L} to near zero and making the total moment predominantly -derived, \mu \approx -2 \mu_B \frac{\mathbf{S}}{\hbar}. This quenching arises because the electrostatic potentials in crystals break the spherical symmetry of free atoms, suppressing orbital circulation. In transition metals like iron (Fe), the magnetic moment is largely from unpaired 3d electron spins, with four unpaired electrons in the free atom yielding an effective moment of approximately $2.2 \mu_B per atom in the metallic state due to band formation and partial quenching. In contrast, rare-earth elements exhibit orbital dominance because their 4f electrons are more localized and less affected by crystal fields, leading to significant \mathbf{L} contributions that can exceed spin moments in materials like samarium nitride (SmN). These atomic-scale moments, when collectively aligned in solids, give rise to the macroscopic magnetization \mathbf{M}.

Quantum Mechanical Basis

The quantum mechanical basis of magnetization originates from the behavior of electrons in atoms and solids, governed by fundamental principles that determine their spin and orbital angular momenta. The Pauli exclusion principle dictates that no two electrons can occupy the same quantum state, which influences the configuration of electrons in atomic orbitals and leads to unpaired spins in partially filled shells. Hund's rules provide a framework for maximizing the total spin angular momentum S in the ground state of multi-electron atoms by filling orbitals with parallel spins before pairing, thereby minimizing electron-electron repulsion and enhancing the magnetic moment. For instance, in transition metal atoms like iron, this results in high-spin configurations with significant net spin, contributing to atomic magnetic moments on the order of several Bohr magnetons. At the level of interacting atoms, the arises from the quantum overlap of wavefunctions, favoring aligned or anti-aligned spins depending on the sign of the . This is modeled by the , H = -J \sum_{\langle i,j \rangle} \mathbf{S}_i \cdot \mathbf{S}_j, where J > 0 promotes ferromagnetic ordering (parallel spins) and J < 0 leads to antiferromagnetic ordering (anti-parallel spins). In ferromagnetic materials like iron, the positive J stabilizes collective spin alignment below the Curie temperature, while in antiferromagnets like manganese oxide, negative J results in staggered spin configurations that cancel macroscopic magnetization. This interaction, rooted in the antisymmetry of the fermionic wavefunction, is the primary mechanism for long-range magnetic order in solids. In crystalline environments, the crystal field generated by surrounding ligands or ions splits the degenerate d-orbitals of transition metals, altering the effective magnetic moments. For octahedral complexes, the five d-orbitals split into lower-energy t_{2g} (dxy, dxz, dyz) and higher-energy e_g (d_{z^2}, d_{x^2-y^2}) sets, with the splitting energy \Delta_o determining whether electrons occupy high-spin (weak field, more unpaired spins) or low-spin (strong field, paired spins) states. This quenching of orbital contributions reduces the total moment compared to free ions; for example, in Ni^{2+} octahedral complexes, the high-spin state yields an effective moment of about 2.8–3.2 \mu_B instead of the spin-only 2.83 \mu_B. Such effects are crucial in coordination compounds and ionic magnets, where symmetry dictates the anisotropy of magnetic properties. In solid-state materials, band theory distinguishes between itinerant and localized electron behaviors, shaping paramagnetic responses. Itinerant electrons in metals form delocalized bands, leading to Pauli paramagnetism where spin-up and spin-down bands shift in a magnetic field, yielding a temperature-independent susceptibility proportional to the density of states at the Fermi level. In contrast, insulators host localized moments on atomic sites, as in , where strong electron correlations prevent band formation and preserve atomic-like spins for Curie-Weiss behavior. This dichotomy explains weak paramagnetism in simple metals like copper versus strong responses in rare-earth compounds.

Macroscopic Relations

Connections Between B, H, and M

In magnetic materials, the magnetic flux density \mathbf{B}, the magnetic field strength \mathbf{H}, and the magnetization \mathbf{M} are interrelated through the constitutive relation \mathbf{B} = \mu_0 (\mathbf{H} + \mathbf{M}) in SI units, where \mu_0 = 4\pi \times 10^{-7} H/m is the permeability of free space. This equation decomposes \mathbf{B} into contributions from the external field \mathbf{H} and the material's response \mathbf{M}, with all quantities expressed as vectors in amperes per meter for \mathbf{H} and \mathbf{M}, and tesla for \mathbf{B}. In vacuum, where no material is present, the magnetization vanishes (\mathbf{M} = 0), simplifying the relation to \mathbf{B} = \mu_0 \mathbf{H}. Within materials, \mathbf{M} arises from the alignment of atomic magnetic moments induced by \mathbf{H}, quantifying the deviation from the vacuum case due to bound currents within the material. For linear isotropic media, where the material response is proportional and direction-independent, \mathbf{M} = \chi_m \mathbf{H}, with \chi_m denoting the dimensionless magnetic susceptibility. Substituting this into the constitutive relation yields \mathbf{B} = \mu_0 (1 + \chi_m) \mathbf{H} = \mu \mathbf{H}, where \mu = \mu_0 (1 + \chi_m) is the permeability of the material. Materials with \chi_m > 0 (paramagnets) enhance the field, while \chi_m < 0 (diamagnets) slightly reduce it, though |\chi_m| \ll 1 for both except in ferromagnets. In anisotropic materials, such as certain crystals, the response varies with direction, requiring a tensorial description: M_i = \chi_{ij} H_j, where \chi_{ij} is the magnetic susceptibility tensor and summation over repeated indices j is implied. The constitutive relation then becomes B_i = \mu_0 (H_i + M_i) = \mu_0 ( \delta_{ij} + \chi_{ij} ) H_j, with \delta_{ij} the Kronecker delta, allowing \mathbf{B} to point differently from \mathbf{H} based on the material's symmetry. At interfaces between regions, boundary conditions ensure continuity consistent with Maxwell's equations in the absence of free surface currents. The normal component of \mathbf{B} is continuous across the boundary: B_{1\perp} = B_{2\perp}. The tangential component of \mathbf{H} satisfies \mathbf{\hat{n}} \times (\mathbf{H}_2 - \mathbf{H}_1) = \mathbf{K}_f, where \mathbf{\hat{n}} is the unit normal and \mathbf{K}_f is the free surface current density; if \mathbf{K}_f = 0, then H_{1\parallel} = H_{2\parallel}. These conditions hold regardless of magnetization differences, as \mathbf{M} effects are incorporated into the volume relations rather than surface discontinuities here.

Magnetic Polarization

Magnetization \mathbf{M}(\mathbf{r}) is fundamentally defined as the density of microscopic magnetic dipole moments within a material, expressed microscopically as \mathbf{M}(\mathbf{r}) = \sum_i \boldsymbol{\mu}_i \delta(\mathbf{r} - \mathbf{r}_i), where \boldsymbol{\mu}_i is the magnetic moment of the i-th dipole located at position \mathbf{r}_i. Macroscopically, this is averaged over a small volume V to yield the magnetization vector \mathbf{M} = \frac{1}{V} \sum_i \boldsymbol{\mu}_i, representing the net magnetic dipole moment per unit volume. This average captures the collective alignment of atomic or molecular magnetic moments, such as those arising from electron spins or orbital currents, leading to the material's overall magnetic response. The concept of magnetization bears a strong analogy to electric polarization \mathbf{P} in dielectrics, where \mathbf{P} denotes the net electric dipole moment per unit volume due to aligned electric dipoles under an electric field. In both cases, an external field—\mathbf{E} for electric dipoles and \mathbf{B} for magnetic dipoles—induces or aligns the microscopic moments, resulting in a macroscopic polarization effect that modifies the internal fields within the material. However, the analogies differ in their field impacts: electric polarization typically reduces the internal electric field, while magnetic polarization enhances the internal magnetic field, as reflected in the relation \mathbf{B} = \mu_0 (\mathbf{H} + \mathbf{M}). In non-uniform magnetization, depolarization fields arise from the spatial variation of \mathbf{M}, generating an internal magnetic field \mathbf{H}_\text{int} that opposes the magnetization. This field is quantified by the demagnetization tensor \mathbf{N}, such that \mathbf{H}_\text{int} = -\mathbf{N} \cdot \mathbf{M}, where \mathbf{N} depends on the sample's geometry and is a dimensionless tensor with trace equal to 1. For example, in a uniformly magnetized sphere, \mathbf{N} = \frac{1}{3} \mathbf{I} (with \mathbf{I} the identity tensor), yielding \mathbf{H}_\text{int} = -\frac{1}{3} \mathbf{M}. These fields stem from effective magnetic "charges" with volume density \rho_m = -\nabla \cdot \mathbf{M} and surface density \sigma_m = \mathbf{M} \cdot \hat{\mathbf{n}}, analogous to bound charges in polarized dielectrics. The effective fields in magnetically polarized media receive contributions from both volume and surface distributions of these dipoles. Volume contributions dominate in regions of smoothly varying \mathbf{M}, arising from the curl \nabla \times \mathbf{M} that produces bound volume currents affecting the magnetic field throughout the material. Surface contributions, in contrast, become prominent at boundaries where \mathbf{M} is discontinuous, manifesting as bound surface currents \mathbf{K}_b = \mathbf{M} \times \hat{\mathbf{n}} that generate localized fields, particularly influencing the depolarization effects in finite samples. This distinction is crucial for understanding field nonuniformities in shaped magnetic materials.

Role in Electromagnetism

Magnetization Currents

In magnetic materials, the magnetization \mathbf{M}, which represents the magnetic dipole moment per unit volume arising from the alignment of atomic magnetic moments, gives rise to effective currents known as bound currents. These currents are not free-flowing charges but result from the microscopic orbital and spin motions of electrons within the material. The bound currents contribute to the total magnetic field in a manner analogous to bound charges in electrostatics, influencing the overall magnetic behavior without requiring external conduction. The volume bound current density \mathbf{J_b} is given by the curl of the : \mathbf{J_b} = \nabla \times \mathbf{M} This expression captures the net current due to the spatial variation of \mathbf{M}; in regions of uniform , \nabla \times \mathbf{M} = 0, so \mathbf{J_b} = 0. For non-uniform \mathbf{M}, such as in where the alignment varies, \mathbf{J_b} can form closed loops or vortex-like patterns that resemble circulating currents within the material volume. Additionally, at surfaces or interfaces where \mathbf{M} changes abruptly, a surface bound current density \mathbf{K_b} appears: \mathbf{K_b} = \mathbf{M} \times \hat{\mathbf{n}} where \hat{\mathbf{n}} is the outward unit normal to the surface. This surface current arises from the tangential component of \mathbf{M} at the boundary, effectively modeling the discontinuity in dipole alignment. To separate the effects of these bound currents from externally controllable free currents \mathbf{J_f}, the auxiliary field \mathbf{H} is introduced in Ampère's law in the form \nabla \times \mathbf{H} = \mathbf{J_f}. Here, \mathbf{H} accounts solely for the free currents, while the bound currents are incorporated into the relation \mathbf{B} = \mu_0 (\mathbf{H} + \mathbf{M}), ensuring that the total magnetic field \mathbf{B} reflects contributions from both. This formulation simplifies calculations in materials by isolating controllable sources. A classic example is a long cylinder with uniform axial magnetization \mathbf{M} = M \hat{\mathbf{z}}, which produces no volume bound current (\mathbf{J_b} = 0) but a circumferential surface current \mathbf{K_b} = M \hat{\phi} on the lateral surface, mimicking the current in a solenoid and yielding \mathbf{B} \approx \mu_0 M inside the cylinder. In contrast, for non-uniform magnetization, such as a varying \mathbf{M} in ferromagnetic domains, the volume current \mathbf{J_b} can generate localized fields that drive domain wall motion or hysteresis effects. These bound currents thus play a crucial role in the macroscopic magnetic response of materials.

Integration into Maxwell's Equations

In electromagnetic theory, magnetization \mathbf{M} is incorporated into Maxwell's equations by accounting for the bound currents it induces within materials, which contribute to the total current alongside free currents \mathbf{J}_f. The general Ampère's law in matter accounts for both free and bound currents: \nabla \times \mathbf{B} = \mu_0 \left( \mathbf{J}_f + \nabla \times \mathbf{M} + \frac{\partial \mathbf{P}}{\partial t} \right) + \mu_0 \epsilon_0 \frac{\partial \mathbf{E}}{\partial t}, or equivalently using the electric displacement \mathbf{D} = \epsilon_0 \mathbf{E} + \mathbf{P}, \nabla \times \mathbf{B} = \mu_0 (\mathbf{J}_f + \nabla \times \mathbf{M}) + \frac{\partial \mathbf{D}}{\partial t}, assuming no surface contributions in the differential form. This modification arises because magnetization represents the density of aligned atomic magnetic moments, producing effective currents that influence the magnetic flux density \mathbf{B}. To simplify the equations in materials, the auxiliary field \mathbf{H} is introduced as \mathbf{H} = \frac{\mathbf{B}}{\mu_0} - \mathbf{M}, leading to the revised Ampère's law \nabla \times \mathbf{H} = \mathbf{J}_f + \frac{\partial \mathbf{D}}{\partial t}, where \mathbf{D} = \epsilon_0 \mathbf{E} + \mathbf{P} is the electric displacement field and \mathbf{P} is the electric polarization. Gauss's law for magnetism remains unchanged as \nabla \cdot \mathbf{B} = 0, reflecting the absence of magnetic monopoles, even in magnetized media. These forms separate free charges and currents from bound effects, facilitating the analysis of material responses. In integral form, the modified Ampère's law follows from Stokes' theorem applied to \nabla \times \mathbf{H}, yielding \oint \mathbf{H} \cdot d\mathbf{l} = I_{f,\text{encl}} + \int_S \frac{\partial \mathbf{D}}{\partial t} \cdot d\mathbf{A}, where I_{f,\text{encl}} is the free current through the surface S bounded by the loop. Similarly, the integral Gauss's law is \int_V \nabla \cdot \mathbf{B} \, dV = \oint_S \mathbf{B} \cdot d\mathbf{A} = 0. For alternating current (AC) fields, time-harmonic analysis uses phasor notation, where magnetization becomes frequency-dependent, expressed through the complex magnetic susceptibility \chi(\omega). The effective permeability is then \mu(\omega) = \mu_0 (1 + \chi(\omega)), allowing \mathbf{B} = \mu(\omega) \mathbf{H} in the frequency-domain Maxwell equations, such as \nabla \times \mathbf{E} = -j\omega \mathbf{B}. This complex form captures dispersion and losses in magnetic materials at high frequencies.

Static and Dynamic Behavior

Magnetostatics

In magnetostatics, the behavior of magnetization is analyzed under steady-state conditions where time derivatives vanish, and magnetic fields arise from steady currents and fixed magnetization distributions. The governing equations are derived from Maxwell's equations in the absence of time variation. Specifically, the divergence of the magnetic flux density B is zero, ∇ · B = 0, reflecting the absence of magnetic monopoles. The curl of the magnetic field strength H equals the free current density J_f, ∇ × H = J_f, where J_f represents currents not bound to the material. The constitutive relation linking these fields is B = μ₀ (H + M), where M is the magnetization and μ₀ is the permeability of free space. Equilibrium configurations of magnetization are determined by minimizing the magnetostatic energy stored in the system. For linear media, where magnetization responds proportionally to the applied field, the relevant energy functional is U = \frac{1}{2} \int \mathbf{B} \cdot \mathbf{H} , dV, integrated over all space; this form represents the total stored magnetostatic energy, with the factor of 1/2 arising from the incremental work to establish the field. In uniform applied fields, solutions inside materials assume M = χ H, where χ is the magnetic susceptibility. However, for finite samples, the internal field H is reduced by the demagnetizing field due to surface magnetic "charges," leading to H = H₀ - N · M, where H₀ is the applied field and N is the demagnetization tensor (with components as shape factors, e.g., N = 1/3 for a sphere along any axis, N = 0 for an infinite cylinder along its axis, and N = 1 for a thin disk perpendicular to the field). Solving for M yields M = χ H₀ / (1 + χ N), illustrating how sample geometry influences the effective susceptibility. In ferromagnetic materials, magnetostatics manifests through nonlinear responses captured in the B-H hysteresis loop, which traces the relation between B and H during field cycling. Starting from a demagnetized state, increasing H aligns domains, raising M toward saturation M_s, the maximum magnetization limited by atomic moments (typically ~10^6 A/m for common ferromagnets like iron). At large H, B ≈ μ₀ M_s since H becomes negligible compared to M. Upon reducing H to zero, a remnant magnetization B_r = μ₀ M_r persists due to domain pinning. Reversing H to the coercive field -H_c (often ~10^3 to 10^5 A/m, depending on microstructure) reduces B to zero, requiring further reversal to -H_c to demagnetize. The loop closes symmetrically, with area representing energy loss per cycle, though in pure static conditions, the loop sketches the stable quasi-static states without dynamic losses. High M_s and H_c are desirable for permanent magnets to maintain strong fields against demagnetization.

Time Evolution and Dynamics

The time evolution of magnetization describes how the magnetic moment density \mathbf{M} changes in response to applied fields and internal interactions, primarily through precessional motion and dissipative relaxation processes that drive the system toward equilibrium. In ferromagnetic and paramagnetic materials, individual atomic magnetic moments precess around an effective magnetic field \mathbf{H}_\mathrm{eff}, leading to a collective dynamics of \mathbf{M}. The fundamental equation governing this motion is the , introduced in 1935, which captures the torque-induced precession as \frac{d\mathbf{M}}{dt} = -\gamma \mathbf{M} \times \mathbf{H}_\mathrm{eff}, where \gamma is the gyromagnetic ratio relating angular momentum to magnetic moment. This form neglects energy dissipation, resulting in perpetual precession without decay. To incorporate damping, T. L. Gilbert reformulated the equation in 1955 using a Lagrangian approach, yielding the widely used Landau-Lifshitz-Gilbert (LLG) equation: \frac{d\mathbf{M}}{dt} = -\gamma \mathbf{M} \times \mathbf{H}_\mathrm{eff} + \frac{\alpha}{M_s} \mathbf{M} \times \frac{d\mathbf{M}}{dt}, where \alpha is the dimensionless Gilbert damping parameter (typically $10^{-2} to $10^{-3} in metals), and M_s = |\mathbf{M}| is the saturation magnetization magnitude. The first term describes precession, while the second introduces phenomenological damping that aligns \mathbf{M} with \mathbf{H}_\mathrm{eff} over time, conserving the magnitude of \mathbf{M} while reducing its misalignment. The effective field \mathbf{H}_\mathrm{eff} includes external fields, exchange interactions, magnetocrystalline anisotropy, and static demagnetization effects. Numerical solutions of the LLG equation, often via micromagnetic simulations, are essential for modeling ultrafast magnetization dynamics in applications like spintronics. For a single spin or uniform \mathbf{M} in a static uniform magnetic field \mathbf{B}, the precession occurs at the Larmor frequency \omega = \gamma B, independent of the initial orientation, as derived by Joseph Larmor in 1897 from classical electrodynamics of orbiting charges. This frequency sets the timescale for coherent oscillations, observable in techniques like ferromagnetic resonance, where \omega / 2\pi ranges from GHz for typical fields of 0.1–1 T. In ordered magnetic materials like ferromagnets, collective excitations arise as spin waves, or magnons, representing propagating deviations of local moments from the ground state alignment. Felix Bloch's 1930 theory established the framework for these excitations, treating them as quantized waves in a Heisenberg spin lattice. For long-wavelength magnons in isotropic ferromagnets under a uniform field, the dispersion relation is approximately \omega(\mathbf{k}) \approx \gamma B + D k^2, where \mathbf{k} is the wavevector, D is the exchange stiffness (on the order of 10–100 meV Ų in transition metals), and the quadratic term reflects the energy cost of spatial variations via nearest-neighbor exchange interactions. This parabolic dispersion enables low-energy, tunable propagation, with group velocities v_g = 2 D k / \hbar reaching km/s for microwave frequencies. In paramagnetic systems, where moments are thermally disordered, the dynamics following a perturbation—such as a radiofrequency pulse—are characterized by relaxation toward thermal equilibrium. In , the longitudinal relaxation time T_1 governs the recovery of the component of magnetization parallel to the static field \mathbf{B}_0, while the transverse relaxation time T_2 describes the dephasing of the perpendicular component due to local field inhomogeneities. These times, typically milliseconds to seconds depending on the nucleus and environment, were formalized by Bloembergen, Purcell, and Pound in 1948 through extended with relaxation terms. For electron spins in , analogous times apply but are shorter (nanoseconds to microseconds) due to stronger interactions.

Material Processes

Magnetization Reversal

Magnetization reversal refers to the processes by which the direction of magnetization in a magnetic material is inverted, typically under the influence of an applied magnetic field, electric current, or thermal fluctuations. This phenomenon is central to the operation of magnetic storage devices, sensors, and spintronic applications, where controlled switching between magnetization states enables data writing and reading. In ferromagnetic materials, reversal can occur through various microscopic mechanisms, each dominant in specific material geometries and conditions, such as single-domain nanoparticles or multidomain bulk samples. One primary mechanism is coherent rotation, described by the Stoner-Wohlfarth model for single-domain particles where the entire magnetization vector rotates uniformly without internal domain formation. In this model, applicable to small, uniformly magnetized grains, the energy landscape features an anisotropy barrier that the magnetization must overcome to switch direction. The energy barrier height is given by \Delta E = K V \left(1 - \frac{H}{H_k}\right)^2 where K is the magnetocrystalline anisotropy constant, V is the particle volume, H is the applied field, and H_k = 2K/M_s is the anisotropy field with M_s the saturation magnetization; reversal occurs when the field reduces the barrier sufficiently for thermal activation or deterministic switching. This model predicts a square hysteresis loop for fields aligned with the easy axis and angular dependence of coercivity, validated in fine particle systems like those in magnetic recording media. In multidomain materials, reversal often proceeds via nucleation and propagation of domain walls, where regions of reversed magnetization form and expand by wall motion under an applied field. Nucleation initiates at defects or surface irregularities, creating a reversed domain, after which the wall propagates if the field exceeds the propagation field. The energy per unit area of a 180° Bloch domain wall, balancing exchange and anisotropy energies, is \sigma = 4 \sqrt{A K} with A the exchange stiffness constant; wall mobility arises from the driving force $2 M_s H per unit area, leading to velocity v = \mu H where \mu is the wall mobility, typically on the order of 10-100 m²/(A·s) in soft ferromagnets like permalloy. This process dominates in bulk magnets and thin films, enabling efficient switching in devices like hard disk drives. Current-induced reversal via spin-transfer torque provides a field-free alternative, particularly in nanoscale multilayer structures such as magnetic tunnel junctions. In this mechanism, a spin-polarized current transfers angular momentum to the local magnetization, exerting a torque that destabilizes the initial state and drives switching. The Slonczewski torque term, derived from quantum mechanical scattering in ferromagnet/nonmagnet/ferromagnet trilayers, is \boldsymbol{\tau} = \frac{\hbar}{2e} \frac{J}{M_s t} \boldsymbol{M} \times (\boldsymbol{M} \times \hat{\mathbf{p}}) where \hbar is the reduced Planck's constant, e the electron charge, J the current density, t the free layer thickness, \boldsymbol{M} the magnetization, and \hat{\mathbf{p}} the fixed layer polarization direction; critical current densities for reversal scale with the damping and anisotropy, enabling sub-nanosecond switching in spin-transfer torque random access memory (STT-MRAM). Thermal activation enables reversal over energy barriers at finite temperatures, even below the coercive field, following the Néel-Brown theory for superparamagnetic particles and thermally assisted switching. This statistical approach treats magnetization as undergoing Brownian motion on the energy surface, with the reversal rate determined by the attempt frequency and barrier height. The mean reversal rate is f = f_0 \exp\left(-\frac{\Delta E}{k T}\right) where f_0 \approx 10^9 Hz is the precession frequency, k Boltzmann's constant, and T temperature; for single-domain particles, this predicts superparamagnetic relaxation times \tau = 1/(2f) that limit data retention in high-density storage, with Néel's original formulation for fine grains and Brown's extension incorporating gyromagnetic precession.

Demagnetization Mechanisms

Thermal demagnetization occurs when the temperature of a ferromagnetic material exceeds its Curie temperature T_c, at which point the spontaneous magnetization M approaches zero due to increased thermal entropy disrupting the aligned spin order. In the mean-field approximation, originally proposed by Weiss for ferromagnets, the magnetization near T_c follows M \propto (T_c - T)^{1/2}, reflecting the second-order phase transition where long-range order is lost. This behavior arises from the competition between exchange interactions favoring alignment and thermal agitation randomizing spins, with T_c determined by the material's exchange energy scale. Field-induced demagnetization involves applying reverse magnetic fields of gradually decreasing magnitude to the existing magnetization, exceeding the coercive field H_c in steps to progressively reduce and randomize domain alignment, achieving near-zero remanent magnetization upon final field removal. The coercive field H_c quantifies the material's resistance to this reversal, stemming from domain wall pinning and nucleation barriers in the hysteresis loop. For hard magnetic materials like NdFeB, high H_c values (e.g., >1 T) maintain remanence against demagnetizing fields, but deliberate application of such decreasing fields can effectively nullify the original M. Time-dependent demagnetization manifests as viscous remanence, where the magnetization decays logarithmically over time due to enabling thermally activated overcoming of energy barriers in magnetic domains. This is characterized by magnetic viscosity S = \frac{dM}{d \ln t}, typically negative for decay, as introduced in early studies of aftereffects in permanent magnets. The coefficient S depends on and , often scaling as S \propto k_B T / V, where V is the activation volume, reflecting the rate of barrier hopping in fine-particle systems or multidomain structures. In (AC) fields, demagnetization arises from dissipation through loops and induced currents, which reduce the effective time-averaged magnetization. losses occur as the material s through repeated magnetization-demagnetization paths, with per proportional to the loop area \oint H \, dM, leading to heating and gradual alignment disorder. losses, dominant at higher frequencies, generate opposing fields from induced currents in conductive materials, further attenuating the dynamic M via ; for example, laminated cores minimize these in transformers. Demagnetization factors, accounting for shape-dependent internal fields, exacerbate losses in non-ellipsoidal samples under AC excitation.

References

  1. [1]
    Magnetic Field Strength H - HyperPhysics
    The quantity M in these relationships is called the magnetization of the material. Another commonly used form for the relationship between B and H is B = μmH.
  2. [2]
    2. Classes of Magnetic Materials | College of Science and Engineering
    Oct 14, 2025 · The origin of magnetism lies in the orbital and spin motions of electrons and how the electrons interact with one another.
  3. [3]
    The Feynman Lectures on Physics Vol. II Ch. 37: Magnetic Materials
    When the magnetization is uniform the currents do not really cancel out precisely; that is, the whirling currents of one electron in one atom and the whirling ...<|control11|><|separator|>
  4. [4]
    [PDF] Electromagnetic Fields and Energy - Chapter 9: Magnetization
    May 9, 2022 · By definition, the magnetization density is the number of magnetic moments per unit volume. For a cubic array with spacing s between the ...
  5. [5]
    Magnetization - Richard Fitzpatrick
    $${\bf v}$ ), or the magnetization ${\bf M}$ (the magnetic dipole moment per unit volume), the magnetic intensity ${\bf H}$ has no clear physical meaning ...
  6. [6]
    [PDF] MAGNETIC MOMENT AND MAGNETIZATION - andrew.cmu.ed
    Magnetization and Dipolar Interactions. We now turn to the definition of the magnetization, M. This is the single most important concept in the chapter ...
  7. [7]
    Chapter 6. Magnetostatic Fields in Matter
    Magnetization (M) is the magnetic dipole moment per unit volume. Materials can be paramagnetic, diamagnetic, or ferromagnetic. The H-field is used to calculate ...
  8. [8]
    [PDF] Lecture Notes 19: Magnetic Fields in Matter I - High Energy Physics
    Thus, in a magnetized object with macroscopic magnetization. (magnetic dipole moment per unit volume) ( )r′. Μ. , each volume element dτ′within the volume v ...
  9. [9]
    [PDF] Units for Magnetic Properties
    Gaussian units are the same as cgs emu for magnetostatics; Mx = maxwell, G = gauss, Oe = oersted; Wb = weber, V = volt, s = second, T = tesla, m = meter, A = ...
  10. [10]
    Conversion between CGS or Gaussian to SI magnetic units
    Jan 19, 2025 · Multiply CGS quantity by Factor to obtain SI quantity, CGS·F = SI. ← Divide SI quantity by Factor to obtain CGS quantity, , CGS = SI / F.
  11. [11]
    Magnetization measurements using SQUID with diamond anvil cells ...
    Jul 24, 2025 · In this review, two types of SQUID magnetometers for pressure (P) of up to 30 GPa are reviewed. One is a commercial SQUID magnetometer with a ...AC magnetic susceptibility... · Magnetization measurements... · VI. RESULTS: DC...
  12. [12]
    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.
  13. [13]
    [PDF] Accuracy of the Reported Moment: Sample Shape Effects
    Oct 4, 2010 · Quantum Design's DC magnetometers, including the SQUID VSM, are calibrated against the cylindrical palladium standard sample. This application ...
  14. [14]
    (PDF) Effect of geometry on magnetization distortion in closed-circuit ...
    Aug 7, 2025 · The degree of the distortion increases as the ratio of length to diameter (L/D) of specimen decreases, and the distortion can reduce the magnetization values ...
  15. [15]
    [PDF] Calibration of vibrating-sample magnetometers
    In the slope method, the magnetometer is calibrated from the low field linear slope of the magnetization curve of a sample of high permeability. ' The primary ...Missing: geometry | Show results with:geometry
  16. [16]
    [PDF] spin.pdf - JILA
    The ge and gp in equations (5.3) and (5.4) are the Landé g-factors of the electron and proton. (NIST 2018) ge = −2.002 319 304 362 56(35) , gp = +5.585 694 ...Missing: formula | Show results with:formula
  17. [17]
    The Magnetic Interaction and the Lande' g-factor - HyperPhysics
    Both the orbital and spin angular momenta contribute to the magnetic moment of an atomic electron. . where g is the spin g-factor and has a value of about 2 ...Missing: source | Show results with:source
  18. [18]
    One hundred years of Alfred Landé's g‐factor - Wiley Online Library
    Oct 5, 2021 · Alfred Landé made use of space quantization in his 1921 papers[9, 10] on the anomalous Zeeman effect and of both space quantization and vector ...
  19. [19]
    [PDF] SOLID STATE PHYSICS PART III Magnetic Properties of Solids - MIT
    1.3 Magnetic Moment and Orbital Angular Momentum . ... Let us consider the quenching of the orbital angular momentum which occurs in materials.
  20. [20]
    Magnetic Moment - an overview | ScienceDirect Topics
    The iron atom as part of the crystal has a magnetic moment of 2.2 BM. Cobalt and nickel have magnetic moments of 1.7 BM and 0.6 BM, respectively. Magnetization ...
  21. [21]
    Rare earth elements and other critical elements in the magnetic ...
    Jun 1, 2022 · Incomplete filling of the 4f subgroup causes the occurrence of both orbital and spin magnetic moments. All REEs except scandium, yttrium ...
  22. [22]
    Twisted phase of the orbital-dominant ferromagnet SmN in a GdN ...
    The orbital-dominant magnetization that is then possible within rare-earth systems facilitates the fabrication of entirely new magnetic heterostructures, and ...
  23. [23]
    [PDF] Hund's Rules - SIUC Physics WWW2
    Same spin orientation makes sure that the probability of two electrons occupying the same orbital is zero (Pauli exclusion), thereby minimizing Coulomb ...
  24. [24]
    Quantum Numbers and Atomic Energy Levels - HyperPhysics
    Another effect on the helium energy levels can be associated with the spin quantum number and arises from the symmetry requirements of the Pauli exclusion ...
  25. [25]
    [PDF] 3 Quantum Magnetism 1 3.5 1D & 2D Heisenberg magnets ...
    9 This is the so-called Heisenberg model. The sign of J can be either antiferromagnetic (J > 0 in this con- vention), or ferromagnetic (J < 0) ...
  26. [26]
    [cond-mat/0511606] Two-dimensional quantum spin-1/2 Heisenberg ...
    Nov 24, 2005 · We study the quantum spin-1/2 Heisenberg model in two dimensions, interacting through a nearest-neighbor antiferromagnetic exchange (J) and a ...
  27. [27]
    [PDF] Superexchange interaction - bingweb
    May 28, 2009 · It is responsible for the Weiss molecular field generated in ferromagnets. This form of the exchange interaction is called a Heisenberg model ...
  28. [28]
    Crystal Field Theory
    The difference between the energy levels in an octahedral complex is called the crystal field splitting energy (Δ o), whose magnitude depends on the charge on ...
  29. [29]
    [PDF] Crystal field and spin Hamiltonian - bingweb
    In studying the magnetic properties of any ionic solid, it is important to know what kind of magnetic ions are present and how they are situated in the lattice.
  30. [30]
    [PDF] arXiv:cond-mat/0006283v1 [cond-mat.str-el] 19 Jun 2000
    In scenario I, suggested by Slater [1] and Stoner [2], band theory alone was proposed to account for itinerant fer- romagnetism: due to the Pauli principle, ...
  31. [31]
    [PDF] electrical and magnetic properties of metals - GovInfo
    critical magnetic fields for over. 800 alloys and compounds. Critical ... maxwell —the unit of magnetic flux in the cgs electromagnetic system. One ...
  32. [32]
    Introduction to Electrodynamics | Cambridge Aspire website
    Aug 17, 2019 · The Fourth Edition provides a rigorous, yet clear and accessible treatment of the fundamentals of electromagnetic theory and offers a sound ...
  33. [33]
    [PDF] Magnetic properties of crystals - HAL
    May 14, 2023 · ∂Hj. Moreover, χ = ∂M. ∂H indicates a tensorial nature of χ since we may write the individual components as: χij = ∂Mi. ∂Hj. Paramagnets ...
  34. [34]
  35. [35]
    [PDF] POLARIZATION AND MAGNETIZATION - UT Physics
    Thus, the magnetic field of a long thin bar magnet looks just like the electric field of a physical dipole, with a positive charge at the North pole and the ...
  36. [36]
    [PDF] Demagnetization Fields 1 The magnetic field inside a magnet
    They oppose each other, hence, the demagnetization field tends to reduce the internal effect of the applied field. It seems to prevent the applied field from ...
  37. [37]
    [PDF] Magnetization, Bound Currents, and the H Field - UT Physics
    The currents (10) and (11) are called the bound currents, by analogy with the bound charges in a dielectric. Example: A uniformly magnetized cylinder, ...
  38. [38]
    [PDF] Magnetic Fields in Matter B
    We change the magnetic dipole to a magnetic dipole density (or magnetization) and then into a bound current J_b and bound surface current K_b using the similar ...
  39. [39]
    None
    ### Key Points on Magnetization Bound Currents
  40. [40]
    [PDF] Section 1: Maxwell Equations
    This polarization current JP has nothing to do with the bound current JM associated with magnetization of the material. The latter involves the spin and orbital ...
  41. [41]
    [PDF] PHY481 - Lecture 29: Magnetic materials
    The magnetization used in Maxwell's equations are the density of aligned moments m 0. Response of electron spins and electron orbits to an applied magnetic ...
  42. [42]
    [PDF] 17 Magnetization current, Maxwell's equations in material media
    proton and neu- trons — produce magnetization currents ∇ × M. – Even bare electrons can produce magnetization ...
  43. [43]
    [PDF] Lecture 2 Review of Maxwell's Equations, EM Energy and Power
    • magnetization and complex permeability. ˆ. ( , , , ). ( , , )cos( ). ˆ ... Describe with one sentence the physical meaning of terms 1, 2, 3,. 4, 5, 6 and 7 in ...<|separator|>
  44. [44]
    Ferromagnetism - Richard Fitzpatrick
    This phenomenon is known as hysteresis. ... . Thus, it is vitally important that a permanent magnet should possess both a large remanence and a large coercivity.
  45. [45]
    [PDF] Lecture 25 - MIT OpenCourseWare
    Derivation of hysteresis loop for a single domain ferromagnet. 5. Coercive field vs. saturation magnetization. Questions you should be able to answer by the ...
  46. [46]
    [PDF] Reprinted from Phys. Zeitsch. der Sow. 8, pp. 153–169 (1935)
    He constructed the domain theory for ferromagnetics (1936) and deduced the equation of motion for a magnetic moment (the. Landau–Lifshits equation). E.M. ...
  47. [47]
    The Landau-Lifshitz equation: 80 years of history, advances, and ...
    Sep 1, 2015 · ... On the theory of the dispersion of magnetic permeability in ferromagnetic bodies,” by L. D. Landau and E. M. Lifshitz was published (Phys.
  48. [48]
    Zur Theorie des Ferromagnetismus - SpringerLink
    Download PDF · Download PDF. Published: March 1930. Zur Theorie des Ferromagnetismus. F. Bloch. Zeitschrift für Physik volume 61, pages 206–219 (1930)Cite this ...
  49. [49]
    A mechanism of magnetic hysteresis in heterogeneous alloys
    ElBidweihy H Rotational Magnetization Lag-Angle Plots Using the Anisotropic Stoner–Wohlfarth Model, IEEE Transactions on Magnetics, 10.1109/TMAG ...
  50. [50]
    [PDF] MEAN-FIELD APPROACH TO MAGNETISM
    Above the critical temperature the system is paramagnetic and no spontaneous magnetization is possible. Mean-field approximation leads thus for an elegant ...
  51. [51]
    [PDF] 2 The Curie–Weiss Model
    In this chapter, we consider its mean-field approximation, in the form of the Curie–Weiss model. Although it is an over-simplification of the Ising model, the ...
  52. [52]
    H
    Weiss Mean Field (or Molecular Field) Theory of Ferromagnetism. This is a prototype for mean field theories and second order phase transitions. In 1907 ...
  53. [53]
    The Hysteresis Loop - Magnetism - NDE-Ed.org
    Coercive Force - The amount of reverse magnetic field which must be applied to a magnetic material to make the magnetic flux return to zero. (The value of H ...
  54. [54]
    Coercive Field - an overview | ScienceDirect Topics
    The coercive field measures the order of magnitude of the fields that must be applied to a material in order to reverse its magnetization.
  55. [55]
    Magnetic Viscosity - NASA ADS
    Under the name of magnetic aftereffect (or time-lag in magnetization or magnetic viscosity) everything is included that is connected with the influence of time ...
  56. [56]
    Magnetic Viscosity - ResearchGate
    Magnetic viscosity also called magnetic aftereffect describes the slow relaxation of magnetization of magnetized bodies. In many cases the magnetization ...
  57. [57]
    Time and temperature dependent magnetic viscosity experiments on ...
    Apr 24, 2023 · The viscosity S ¼ dM(t)/dln(t), determined as the slope of the M vs ln(t) curve, is given by. 2MSkBTf(E), where MS is the saturation ...
  58. [58]
    Hysteresis Loss - an overview | ScienceDirect Topics
    Hysteresis loss is defined as the energy loss that occurs due to the oscillation of magnetic dipoles in magnetic materials, such as steel and nickel, ...
  59. [59]
    Hysteresis Loss and Eddy Current Loss: What's the Difference?
    Mar 30, 2018 · Hysteresis loss is caused by the magnetization and demagnetization of the core as current flows in the forward and reverse directions.
  60. [60]
    [PDF] Internal fields in magnetic materials and superconductors
    Demagnetization factors, initial susceptibili- ties and hysteresis losses are discussed. Keywords: demagnetization factors; hysteresis; internal fields; ...