Fact-checked by Grok 2 weeks ago

Introduction to Solid State Physics

is the branch of physics that investigates the physical properties of solid materials, particularly crystalline solids, through the lens of , , and to explain phenomena such as , thermal behavior, , and optical responses at the and levels. It focuses on how the ordered arrangement of atoms in influences macroscopic properties, distinguishing solids from other states of matter by their rigidity and fixed volume. This field, also overlapping with , emphasizes the role of interactions, lattice vibrations, and defects in determining material characteristics. Central to solid state physics are concepts like crystal structures and , which describe the periodic atomic arrangements in solids such as face-centered cubic (fcc), body-centered cubic (bcc), and hexagonal close-packed . techniques, including , , and , are used to probe these structures experimentally. , involving phonons as quantized vibrational modes, account for like specific and elasticity, while the provides an initial framework for understanding metallic conduction before more advanced band theory. Electronic band structure is a cornerstone, classifying materials as metals, semiconductors, insulators, or semimetals based on energy gaps and the filling of bands near the . In semiconductors, doping introduces carriers that enable applications in , while in metals, the and Sommerfeld models explain transport properties like resistivity and the . Magnetic properties arise from electron spins and orbital motions, leading to , , , and phenomena like the in superconductors. Optical and thermal responses, including excitons and , further illustrate how quantum effects manifest in solids. The field has profound implications for technology, underpinning developments in semiconductors for , superconductors for transmission, and for advanced devices, with ongoing research exploring low-dimensional systems like and quantum dots. Experimental tools such as , scanning tunneling microscopy, and continue to refine theoretical models.

Overview and Scope

Definition and Key Concepts

Solid-state physics is the study of the properties of solid materials and how these properties emerge from the interactions of their constituent atoms. It employs principles from to describe behavior, to analyze atomic arrangements, and to understand fields and responses in solids, thereby linking microscopic structures to macroscopic phenomena such as electrical and . This field is distinct from , which broadly encompasses the study of both solids and liquids (as well as other dense phases like plasmas), focusing on collective behaviors in densely packed matter regardless of rigidity. In contrast to , which emphasizes the synthesis, chemical bonding mechanisms, and reactivity of solids through molecular orbital and valence concepts, solid-state physics prioritizes the physical properties arising from electronic structure and lattice dynamics, often using computational methods to predict band structures and transport. At its core, revolves around the atomic-scale organization of matter in solids. Crystalline solids, such as (NaCl), feature atoms arranged in a periodic with long-range translational order, where repeating units form a three-dimensional pattern that dictates and properties like planes. Amorphous solids, exemplified by , lack this periodicity, exhibiting short-range order but no extended , leading to isotropic behavior and gradual softening rather than sharp points. In these structures, positively charged ions typically form the rigid framework, while valence electrons mediate bonding—either localized in insulators or delocalized in conductors—fundamentally influencing thermal, electrical, and magnetic responses. A foundational prerequisite is the quantum description of electrons in isolated atoms, where each electron occupies discrete energy states defined by four quantum numbers: the principal quantum number n (determining energy level), orbital angular momentum l (shape), magnetic m_l (orientation), and spin m_s (\pm 1/2). The Pauli exclusion principle ensures no two electrons share identical quantum numbers, filling shells from lowest to highest energy to form stable atomic configurations essential for understanding how these states broaden into bands in solids.

Historical Context and Importance

The foundations of solid state physics were laid in the early through pioneering work on structures and applied to solids. In 1913, and William Lawrence Bragg developed X-ray diffraction techniques, enabling the determination of atomic arrangements in via , which revolutionized the study of material structures. This breakthrough provided experimental tools to probe the periodic nature of solids, bridging classical crystallography with emerging quantum ideas. Building on this, Felix Bloch's 1928 thesis introduced the concept of electron waves in periodic potentials, demonstrating how could explain electron behavior in through Bloch waves, a cornerstone for understanding electronic properties in solids. The field experienced explosive growth after , fueled by technological demands and fundamental advances. In December 1947, , Walter Brattain, and at Bell Laboratories invented the , a that amplified signals and replaced bulky vacuum tubes, laying the groundwork for modern electronics. That same year, the established its Division of Solid State Physics (DSSP), formalizing the discipline and reflecting its rising prominence amid wartime research on materials. By the 1950s and 1960s, influential textbooks like Charles Kittel's Introduction to Solid State Physics (first edition, 1953) and Neil Ashcroft and N. David Mermin's Solid State Physics (1976) synthesized these developments, providing accessible frameworks for quantum treatments of lattice vibrations, band structures, and . In 1978, the renamed the DSSP to the to encompass broader studies of quantum many-body systems, including liquids and , beyond just solids, amid evolving research on phase transitions and . This shift underscored the field's maturation into a central pillar of physics, with profound technological impacts: solid state principles enabled the , powering computers, light-emitting diodes (LEDs), and integrated circuits that transformed computing and communications. By 2023, the global market reached $533 billion in revenue, highlighting its economic scale and role in driving digital innovation. Post-2010, the discipline has increasingly focused on , such as topological insulators and graphene-based systems, where emergent phenomena like protected edge states promise advances in and .

Fundamental Crystal Structures

Bravais Lattices and Unit Cells

In three-dimensional space, crystal lattices can be classified into 14 distinct Bravais lattices, named after Auguste Bravais who systematically enumerated them based on translational symmetry and point group operations. These lattices are grouped into seven crystal systems—triclinic, monoclinic, orthorhombic, tetragonal, trigonal (or rhombohedral), hexagonal, and cubic—each characterized by specific symmetry constraints on the lattice parameters a, b, c (edge lengths) and angles \alpha, \beta, \gamma. For instance, the cubic system requires a = b = c and \alpha = \beta = \gamma = 90^\circ, while the triclinic system has no such restrictions. A Bravais lattice is defined as an infinite array of discrete points where each point has an identical environment, generated by integer combinations of three basis vectors \mathbf{a}, \mathbf{b}, \mathbf{c}. The fundamental building block of a Bravais lattice is the unit cell, which is the smallest volume containing all lattice points when translated. A primitive unit cell has lattice points only at its corners and a volume equal to the lattice's primitive volume V = |\mathbf{a} \cdot (\mathbf{b} \times \mathbf{c})|, ensuring one lattice point per cell. In contrast, a conventional unit cell may be larger to better reflect the lattice's symmetry, incorporating additional lattice points at face centers or body centers; for example, the conventional cubic unit cell for a face-centered cubic (FCC) lattice includes four lattice points. This distinction allows for clearer visualization and calculation of properties while preserving the lattice's periodicity. The provides a in momentum space, essential for understanding and wave propagation in . Defined as the set of all vectors \mathbf{G} such that \mathbf{G} \cdot \mathbf{R} = 2\pi n for any direct lattice vector \mathbf{R} = m\mathbf{a} + n\mathbf{b} + p\mathbf{c} (with integers m, n, p) and integer n, the reciprocal basis vectors are \mathbf{a}^* = 2\pi \frac{\mathbf{b} \times \mathbf{c}}{\mathbf{a} \cdot (\mathbf{b} \times \mathbf{c})}, and cyclically for \mathbf{b}^*, \mathbf{c}^*. The is the Fourier transform of the direct lattice, transforming periodic density in real space to discrete points in reciprocal space, as originally introduced by Paul Peter Ewald. Reciprocal lattice vectors take the form \mathbf{G}_{hkl} = 2\pi (h \mathbf{a}^* + k \mathbf{b}^* + l \mathbf{c}^*), where h, k, l are integers labeling the vector. Miller indices (hkl) denote families of parallel lattice planes in a crystal, introduced by William Hallowes Miller as a systematic notation for crystallographic directions and orientations. To assign indices, reciprocals of the intercepts of a plane with the lattice axes (in units of a, b, c) are taken and reduced to smallest integers; planes parallel to an axis have infinite intercept, yielding zero index (e.g., (100) for planes parallel to b and c). Negative indices use overbar notation, like (\bar{1}00). The interplanar spacing d_{hkl} between consecutive (hkl) planes is given by d_{hkl} = \frac{2\pi}{|\mathbf{G}_{hkl}|}, linking geometry directly to reciprocal space for diffraction analysis. Representative examples illustrate these concepts. The simple cubic , with primitive edges a = b = c at right angles and one atom per , is exemplified by (Po), the only elemental metal adopting this structure at ambient conditions due to relativistic effects stabilizing the low-coordination geometry. The face-centered cubic (FCC) adds atoms at face centers to the cubic primitive , yielding a conventional with four atoms and high packing (\approx [74](/page/74)\%); (Cu) adopts this structure, with parameter a \approx 0.3615 nm at . The body-centered cubic (BCC) places an additional atom at the body center, resulting in two atoms per conventional and packing \approx 68\%; alpha-iron (Fe) exhibits this structure below 912°C, with a \approx 0.2866 nm.

Crystal Symmetries and Defects

Crystal symmetries arise from the periodic arrangement of atoms in a lattice, which can be described by point groups and space groups that capture the rotational, reflectional, and translational invariances of the structure. Point groups classify the symmetry operations that leave a point fixed, such as rotations and reflections, and there are 32 distinct crystallographic point groups in three dimensions, arising from the crystallographic restriction theorem that limits possible rotation axes to 1, 2, 3, 4, or 6-fold symmetries. Space groups extend this by incorporating translations via screw axes and glide planes, resulting in 230 unique space groups that fully describe the symmetries of periodic crystals. For example, the cubic crystal system, common in materials like sodium chloride, belongs to the point group O_h, which includes 48 symmetry operations such as 3-fold rotations along body diagonals, 4-fold rotations along face normals, and inversion through the center, enabling isotropic properties in high-symmetry directions. In real crystals, deviations from perfect periodicity introduce defects that disrupt these symmetries and profoundly influence material properties. Point defects are zero-dimensional imperfections, including vacancies (missing atoms) and interstitials (extra atoms squeezed into non-lattice sites), which occur due to or processing conditions. The concentration of vacancies follows c_v = \exp(-E_f / [kT](/page/KT)), where E_f is the formation energy (typically 1-2 for metals) and kT is , leading to about one vacancy per million sites at in . In ionic crystals like NaCl, Schottky defects maintain charge neutrality by creating equal numbers of cation and anion vacancies, with pair formation energies around 2-3 eV, resulting in concentrations on the order of 10^{-7} at 800 K. Line defects, or dislocations, are one-dimensional and include edge dislocations (extra half-planes of atoms) and screw dislocations (shear distortions), with typical densities of $10^6 to $10^{12} cm^{-2} in annealed to deformed metals. Plane defects, such as boundaries, arise at interfaces between crystalline regions of different orientations, while stacking faults disrupt layer sequencing in close-packed structures. These defects' formation energies vary: for example, edge dislocations in aluminum have energies of about 0.5 per atomic distance along the line. Defects play crucial roles in transport and mechanical behavior; vacancies facilitate atomic diffusion by providing sites for atoms to jump, enabling self-diffusion coefficients D = a^2 \nu \exp(-(E_m + E_f)/kT), where E_m is migration energy and a is lattice spacing, as seen in metals where vacancy-mediated diffusion dominates at high temperatures. Dislocations enable by allowing slip along crystallographic planes under , which is essential for the of metals; without them, pure crystals would be brittle, but dislocation motion via glide and climb permits large deformations, as exemplified in face-centered cubic metals like where multiply during deformation to accommodate strains up to 50%. Grain boundaries impede dislocation motion, strengthening materials per the Hall-Petch relation, but also serve as paths for . X-ray diffraction serves as a primary probe of crystal symmetries and detects defects through broadened or absent peaks. The Laue conditions for constructive interference require that the scattered wavevector difference equals a reciprocal lattice vector: \Delta \mathbf{k} = \mathbf{k}_f - \mathbf{k}_i = \mathbf{G}, where |\mathbf{k}_i| = |\mathbf{k}_f| = 2\pi / \lambda for elastic scattering, ensuring phase coherence across the lattice planes perpendicular to \mathbf{G}. This leads to Bragg's law, derived by considering the path length difference for rays reflecting off successive planes separated by distance d. For two parallel rays incident at angle \theta to the planes, the extra path length for the second ray is $2d \sin \theta due to the incoming and outgoing segments. Constructive occurs when this difference equals an integer multiple of the : $2d \sin \theta = n \lambda, where n = 1, 2, \dots. $2 d \sin \theta = n \lambda This equation, with d = 2\pi / | \mathbf{G} |, predicts peaks at specific angles, revealing parameters and symmetries; defects like dislocations cause peak broadening via fields, while point defects reduce intensity without shifting positions.

Lattice Dynamics

Vibrational Modes and Phonons

In crystals, atomic vibrations are modeled using the , where are expanded to quadratic order, resulting in a set of coupled oscillators whose modes describe the collective dynamics. This approach, developed in the Born-von Kármán framework, assumes and nearest-neighbor interactions to solve for the . A simple illustration is the one-dimensional diatomic chain, consisting of alternating atoms of masses m_1 and m_2 connected by springs of constant K and spacing a. In the case of equal masses m_1 = m_2 = m, the acoustic branch simplifies to \omega(k) = \sqrt{\frac{2K}{m}} \left| \sin\left(\frac{ka}{2}\right) \right|, where k is the wave vector, revealing a linear relation at long wavelengths (k \to 0) corresponding to sound propagation and a maximum frequency at the Brillouin zone boundary. For unequal masses, the dispersion splits into acoustic and optical branches, with the optical branch exhibiting a frequency gap at k = 0 due to the relative motion of unlike atoms. To incorporate quantum mechanics, the normal modes are quantized by treating the lattice as a field of harmonic oscillators, introducing phonon quasiparticles as bosonic excitations. Each mode with frequency \omega_{\mathbf{q}j} (where \mathbf{q} is the wave vector and j labels the branch) is described by creation a^\dagger_{\mathbf{q}j} and annihilation a_{\mathbf{q}j} operators satisfying [a_{\mathbf{q}j}, a^\dagger_{\mathbf{q}'j'}] = \delta_{\mathbf{qq}'} \delta_{jj'}, with the Hamiltonian for the mode given by \hbar \omega_{\mathbf{q}j} (a^\dagger_{\mathbf{q}j} a_{\mathbf{q}j} + 1/2). In three dimensions, a crystal with N primitive cells and p atoms per cell supports $3Np normal modes, yielding $3p branches: typically three acoustic branches (vanishing frequency at \mathbf{q} = 0) and $3p - 3 optical branches (finite frequency at \mathbf{q} = 0) for p > 1. The distribution of these modes is captured by the phonon density of states g(\omega), which counts the number of modes per interval. In the Debye approximation, valid for low frequencies where dispersion is linear (\omega = v_s |\mathbf{q}|, with v_s the ), the in three dimensions follows g(\omega) \propto \omega^2 up to a cutoff frequency \omega_D, ensuring the total number of modes matches $3N. In face-centered cubic (FCC) lattices, such as those of noble metals like , the three acoustic branches consist of one longitudinal mode (displacements parallel to propagation) and two degenerate transverse modes (displacements perpendicular to propagation), with dispersion relations measured via neutron scattering showing the longitudinal branch having higher velocities and frequencies than the transverse ones due to stronger restoring forces along the propagation direction.

Properties from Phonons

In solids, the arises predominantly from the vibrational of the , modeled as phonons. At high temperatures, the classical Dulong-Petit law predicts that the at constant volume C_V approaches $3R per atom, where R is the , equivalent to $3Nk_B for N atoms and k_B Boltzmann's constant; this reflects the equipartition of energy among three quadratic per atom. This law holds well for many solids above but fails at low temperatures, where C_V decreases faster than expected classically. To address this discrepancy, developed a quantum mechanical model in 1912, treating phonons as a gas of bosons with a linear up to a cutoff frequency \omega_D, leading to C_V \propto T^3 at low temperatures T \ll \Theta_D, where the Debye temperature is defined as \Theta_D = \frac{\hbar \omega_D}{k_B}. The Debye temperature characterizes the temperature scale below which quantum effects freeze out low-frequency modes, with typical values ranging from about 100 K for lead to over 1000 K for diamond, influencing the heat capacity's temperature dependence. In insulators, the phonon contribution dominates the specific heat across most temperatures, whereas in metals, phonons provide the leading term at higher temperatures, underscoring their universal role in lattice thermodynamics. Thermal expansion in solids links to phonon anharmonicity, where lattice vibrations cause volume changes upon heating. The Grüneisen parameter \gamma, introduced by Eduard Grüneisen, quantifies this by relating the relative change in phonon frequencies to volume: \gamma = - \frac{V}{\omega} \frac{d\omega}{dV}, typically around 1-2 for many materials, predicting thermal expansion coefficients proportional to C_V through the relation \alpha = \frac{\gamma C_V}{3V B}, with B the bulk modulus. This parameter bridges microscopic vibrational shifts to macroscopic expansion, explaining why materials expand more at higher temperatures where anharmonic effects intensify. Thermal conductivity \kappa in insulating solids is governed by phonon transport, analogous to gas kinetic theory, given by \kappa = \frac{1}{3} C_V v l, where v is the average phonon speed and l the mean free path limited by scattering processes. At low temperatures, boundary scattering dominates, yielding large l and high \kappa, while at higher temperatures, anharmonic three-phonon interactions, particularly Umklapp processes that conserve crystal momentum only modulo a reciprocal lattice vector, limit l and cause \kappa to decrease. These Umklapp scattering events are crucial for thermal resistance in perfect crystals, enabling heat flow without net momentum transfer to the lattice.

Electronic Structure of Solids

Free Electron Gas Model

The gas model provides the simplest quantum mechanical description of conduction electrons in metals, treating them as a non-interacting ensemble of fermions confined to a volume V of the solid. This approach assumes that the electrons move freely as plane waves in a constant potential, neglecting any periodic variation due to the ionic lattice, and obey Fermi-Dirac statistics at temperature, where the system fills states up to a maximum energy known as the . are imposed on the wavefunctions to simulate an infinite without surface effects, leading to discrete momentum states spaced by \Delta k_x = 2\pi / L in each direction for a cubic volume L^3 = V. In this model, the ground state configuration forms a filled Fermi sphere in momentum space, with radius k_F = (3\pi^2 n)^{1/3}, where n = N/V is the electron density. The corresponding Fermi energy is given by E_F = \frac{\hbar^2}{2m} (3\pi^2 n)^{2/3}, determining the energy scale for electronic excitations at low temperatures. The density of states per unit volume for these free electrons, which counts the number of available states per energy interval, follows g(E) \propto \sqrt{E}, specifically g(E) = \frac{3}{2} \frac{n}{E_F} \sqrt{\frac{E}{E_F}} near the Fermi level, reflecting the quadratic dispersion E = \frac{\hbar^2 k^2}{2m}. At finite but low temperatures, thermal excitations are limited to states near E_F due to the , leading to the Sommerfeld expansion for thermodynamic properties. The electronic per unit volume is linear in , C_{el} = \gamma T, where the \gamma = \frac{\pi^2}{3} k_B^2 g(E_F) captures the enhanced effective number of excitable electrons compared to classical predictions. The model successfully explains several observed properties of simple metals. For electrical conductivity, it yields the Drude-like expression \sigma = \frac{n e^2 \tau}{m} using the Fermi velocity v_F = \sqrt{2 E_F / m} for , providing a quantum foundation that agrees well with experiment for metals when relaxation time \tau is treated semiclassically. Additionally, it predicts Pauli paramagnetism, arising from the spin alignment of electrons near E_F in a , with \chi = \mu_B^2 g(E_F), which matches measurements in non-magnetic metals.

Nearly Free Electron and Tight-Binding Models

The Bloch theorem provides the foundational framework for understanding wavefunctions in a periodic potential. It asserts that the solutions to the in a periodic take the form \psi_{\mathbf{k}}(\mathbf{r}) = e^{i\mathbf{k}\cdot\mathbf{r}} u_{\mathbf{k}}(\mathbf{r}), where u_{\mathbf{k}}(\mathbf{r}) is a periodic function with the same periodicity as the , and \mathbf{k} is a wavevector within the first Brillouin zone. This form ensures that the wavefunction satisfies the translational symmetry of the , allowing to propagate as plane waves modulated by the periodicity, as derived by Felix Bloch in his analysis of motion in lattices. The extends the gas by incorporating a weak periodic potential from the ionic , treated as a . In this approach, unperturbed plane-wave states are degenerate at the boundaries, where the periodic potential mixes states with wavevectors \mathbf{k} and \mathbf{k} + \mathbf{G} (with \mathbf{G} a vector), opening energy gaps. The gap size at these boundaries is given by \Delta E = 2|V_{\mathbf{G}}|, where V_{\mathbf{G}} is the component of the potential corresponding to \mathbf{G}. This , building on Bloch's framework, explains the formation of energy bands and gaps in metals and semiconductors with weak scattering, such as alkali metals. In contrast, the tight-binding model starts from localized orbitals on sites, assuming strong binding to individual atoms with weak overlap between neighbors. The wavefunction is constructed as a of these orbitals, leading to Bloch states where the arises from hopping between sites, characterized by the hopping t. For a simple cubic with z nearest neighbors, the is W = 2z|t|. In a one-dimensional chain with a, the is E(k) = -2t \cos(ka), illustrating how the periodic overlap produces band structure with width $4|t|. This model, formalized through the method, is particularly effective for insulators and semiconductors where electrons are tightly bound, such as in covalent solids like . Brillouin zones are constructed in reciprocal space as the Wigner-Seitz cells centered on reciprocal lattice points, defining the unique range of wavevectors \mathbf{k} for Bloch states. The first is the primitive cell bounded by planes perpendicular to reciprocal lattice vectors and bisecting them, ensuring no overlap in the reduced zone scheme where all bands are folded back into this zone. Higher zones extend this construction outward, with boundaries corresponding to Bragg reflection conditions that cause band folding. This zoning scheme, essential for visualizing Fermi surfaces and band structures, directly follows from the periodicity imposed by and the reciprocal lattice geometry.

Band Theory and Semiconductors

Energy Bands and Gaps

In , the electronic structure of crystalline solids arises from the overlap of atomic orbitals as atoms are brought together in a periodic , leading to the formation of continuous bands rather than discrete levels. These bands represent allowed ranges for electrons, described by Bloch states that account for the periodic potential of the crystal. The valence band consists of states typically filled by valence electrons at , while the conduction band comprises higher- states that are empty in insulators and semiconductors but partially occupied in metals. Between these bands lies the band gap, a forbidden energy range E_g where no electron states exist, determining the material's electrical properties. Band gaps are classified as direct or indirect based on the crystal momentum k: in direct gaps, the conduction band minimum and valence band maximum occur at the same k-point, allowing efficient optical transitions; in indirect gaps, they occur at different k-points, requiring phonon assistance for momentum conservation. Solids are thus classified by their band structures: metals exhibit overlapping valence and conduction bands with no gap, enabling high conductivity, as in where the $4s band overlaps the filled $3d band. Insulators have large band gaps exceeding 5 , such as with E_g \approx 5.5 , preventing electron excitation at . Semiconductors feature smaller gaps around 1 , like with an indirect E_g \approx 1.12 , allowing thermal excitation across the gap. The Fermi level E_F, the highest occupied energy at absolute zero, positions differently across material classes: it lies within the overlapping bands of metals, facilitating free electron movement; in insulators and semiconductors, it resides in the band gap, with the valence band fully occupied below it. In intrinsic semiconductors, E_F remains approximately fixed near the band gap center, showing weak temperature dependence due to equal electron and hole populations. Within energy bands, the g(E) quantifies available states per energy interval, varying with band curvature via g(E) \propto \sqrt{E} in three dimensions for free-like electrons but exhibiting singularities at critical points. Van Hove singularities occur where the band dispersion \epsilon(k) has saddle points or extrema (\nabla_k \epsilon = 0), causing g(E) to diverge or exhibit sharp peaks, influencing properties like electronic specific heat and instabilities in solids.

Intrinsic and Extrinsic Semiconductors

In intrinsic semiconductors, charge carriers are generated solely through thermal excitation of electrons from the valence band to the conduction band across the energy bandgap E_g, resulting in equal concentrations of electrons (n) and holes (p) given by the intrinsic carrier concentration n_i. The expression for n_i is derived from Fermi-Dirac statistics and the in the bands, yielding n_i = \sqrt{N_c N_v} \, e^{-E_g / 2kT}, where N_c and N_v are the effective densities of states in the conduction and valence bands, respectively, k is Boltzmann's constant, and T is the temperature. This concentration increases exponentially with temperature, reflecting the dominance of in pure semiconductors like or without impurities. Extrinsic semiconductors are created by intentionally introducing impurities, or dopants, to control carrier concentrations and type, significantly altering electrical properties compared to the intrinsic case. In n-type semiconductors, donor impurities such as (P) in provide extra electrons; these donors occupy substitutional lattice sites and form shallow levels approximately 0.045 below the conduction edge, from which electrons are easily thermally excited at . Conversely, p-type semiconductors incorporate acceptor impurities like (B) in , which create shallow levels about 0.045 above the valence edge, accepting electrons from the valence band and generating holes as majority carriers. Under charge neutrality and assuming complete of dopants, the electron concentration n satisfies n = p + N_d - N_a, where N_d and N_a are the donor and acceptor concentrations, respectively. A key relation in both intrinsic and extrinsic semiconductors is the , np = n_i^2, which holds at regardless of doping and stems from the balance of and recombination processes. Doping shifts the E_F: in n-type materials, E_F moves closer to the conduction edge due to the increased , while in p-type, it approaches the valence edge. The of extrinsic semiconductors varies across temperature regimes: at low temperatures (freeze-out), carriers are bound to levels, yielding low ; at intermediate temperatures (extrinsic regime), ionized dopants provide a nearly constant majority carrier concentration; and at high temperatures (intrinsic regime), thermal dominates, reverting to that of an intrinsic material.

Transport Phenomena

Electrical Conductivity and Drude Model

Electrical conductivity in solids, particularly metals, arises from the motion of charge carriers, primarily electrons, under an applied . In the classical picture, these electrons behave like a gas of particles that drift in response to the field but undergo frequent collisions with lattice ions and impurities, leading to a finite conductivity. The provides the foundational classical description of this process, treating electrons as free particles scattered randomly with a characteristic mean free time \tau. Proposed by Paul Drude in 1900, the model assumes that conduction electrons in metals have a density n comparable to the number of valence electrons per atom, as estimated from the free electron gas model. In the presence of an electric field \mathbf{E}, each electron acquires a drift velocity \mathbf{v}_d = -\frac{e \tau}{m} \mathbf{E}, where e is the electron charge and m is its mass, after averaging over collisions that randomize velocities every \tau. The resulting current density is \mathbf{J} = -n e \mathbf{v}_d = \frac{n e^2 \tau}{m} \mathbf{E}, yielding the DC conductivity \sigma = \frac{n e^2 \tau}{m}. The electron mobility, defined as \mu = \frac{e \tau}{m}, quantifies the ease of carrier drift, with \sigma = n e \mu. This formula successfully predicts the order of magnitude of conductivity in many metals at room temperature, where \tau \approx 10^{-14} s. The DC resistivity is the reciprocal, \rho = 1/\sigma = \frac{m}{n e^2 \tau}, which increases with temperature due to enhanced scattering from lattice vibrations, making \tau decrease. Matthiessen's rule, empirically established in the 1860s, states that the total resistivity decomposes additively into temperature-dependent thermal scattering \rho_T(T) and temperature-independent impurity scattering \rho_i: \rho = \rho_T(T) + \rho_i. This separation holds well for dilute alloys and pure metals, allowing isolation of intrinsic lattice contributions from defect effects. For example, in , \rho_i dominates at cryogenic temperatures, while \rho_T is linear in T near . For alternating fields, the Drude model extends to AC conductivity via a frequency-dependent response. The complex dielectric function is \epsilon(\omega) = 1 - \frac{\omega_p^2}{\omega(\omega + i/\tau)}, where the plasma frequency \omega_p = \sqrt{\frac{4\pi n e^2}{m}} characterizes collective electron oscillations, typically in the ultraviolet for metals like sodium (\omega_p \approx 8.8 \times 10^{15} rad/s). At low frequencies (\omega \ll 1/\tau), this recovers the DC limit, but at high frequencies, it predicts metallic reflection below \omega_p and transparency above, aligning with observed optical properties. Drude derived this in his 1900 extension, incorporating damped harmonic motion for electrons. Despite its successes, the has key limitations, notably at low temperatures. It predicts \rho \to 0 as T \to 0 since \tau \to \infty without , yet experiments show a finite residual resistivity \rho_0 from impurities and defects, unexplained classically. This discrepancy, evident in pure metals like where \rho plateaus below 20 K, highlights the need for quantum treatments of .

Thermoelectric and Hall Effects

The describes the generation of a transverse voltage across a or when subjected to a perpendicular to an applied , resulting from the deflecting s. This phenomenon, first observed by Edwin Hall in 1879 using thin gold foil, provides a direct probe of properties in solids. The \mathbf{F} = q (\mathbf{v} \times \mathbf{B}) causes electrons (or holes) to accumulate on one side of the sample, establishing an E_y that balances the magnetic deflection in . The Hall coefficient R_H, defined as R_H = \frac{E_y}{j_x B_z}, quantifies this and equals R_H = -\frac{1}{n e} for electrons in single-carrier systems, where n is the carrier density and e is the ; the negative sign distinguishes electron conduction from positive R_H = \frac{1}{p e} for . This measurement reveals both the type ( or ) and density of majority , independent of mechanisms, as the deflection depends only on carrier and strengths. In practice, Hall sensors exploit this for non-contact detection in devices like position encoders and current monitors, achieving sensitivities down to microtesla levels. Thermoelectric effects couple electrical and thermal transport in solids, enabling conversion between heat and electricity without moving parts. The Seebeck effect, discovered by Thomas Seebeck in 1821, generates a voltage across a \Delta T, characterized by the S = -\frac{\Delta V}{\Delta T}, typically on the order of 10–1000 \muV/K in semiconductors. The related Peltier effect, identified by Jean Peltier in 1834, produces heat absorption or release at junctions of dissimilar materials when current flows, with the Peltier coefficient \Pi = S T linking the two via the Kelvin relation. These effects underpin solid-state and power generation, as in thermoelectric generators using waste heat. The efficiency of thermoelectric materials is assessed by the dimensionless figure of merit ZT = \frac{S^2 \sigma T}{\kappa}, where \sigma is electrical , T is absolute , and \kappa is thermal conductivity; values exceeding 1 indicate practical viability, with optimization requiring high S and \sigma alongside low \kappa. Thermoelectric coolers, based on the Peltier effect, are widely used in cooling and portable devices, offering silent operation and precise temperature control up to \Delta T \approx 70 K. These are theoretically described by the semiclassical Boltzmann transport equation, which governs the evolution of the distribution function f(\mathbf{r}, \mathbf{k}, t) under electric \mathbf{E}, magnetic \mathbf{B}, and gradients: \frac{\partial f}{\partial t} + \mathbf{v} \cdot \nabla_{\mathbf{r}} f + \frac{\mathbf{F}}{\hbar} \cdot \nabla_{\mathbf{k}} f = \left( \frac{\partial f}{\partial t} \right)_{\text{[coll](/page/Coll)}}. In the , the collision term simplifies to -\frac{f - f_0}{\tau}, where f_0 is the equilibrium Fermi-Dirac distribution and \tau is the momentum relaxation time, allowing computation of coefficients like R_H, S, and \sigma via integrals over the without solving the full collision integral. This approach builds on the by incorporating band structure and field perturbations. For materials with mixed and carriers, such as intrinsic semiconductors, a two-band model accounts for contributions from both conduction and valence bands, yielding an effective Hall R_H = \frac{p \mu_h^2 - n \mu_e^2}{(p \mu_h + n \mu_e)^2 e} and modified S = \frac{S_e \sigma_e + S_h \sigma_h}{\sigma_e + \sigma_h}, where subscripts denote carrier type and \mu is . This framework is essential for analyzing conduction in thermoelectric alloys like Bi_2Te_3, enabling optimization of ZT by tuning doping to suppress minority carrier effects.

Magnetic and Optical Properties

Diamagnetism, Paramagnetism, and Ferromagnetism

arises from the induced magnetic moments in atoms or ions when exposed to an external , resulting in a weak repulsion and negative . In solids, this effect is described by Larmor diamagnetism, where the orbital motion of electrons precesses around the field direction, generating an opposing moment. The diamagnetic for a system of n electrons per unit volume is given by \chi = -\frac{\mu_0 n e^2 \langle r^2 \rangle}{6m}, where \mu_0 is the vacuum permeability, e and m are the electron charge and mass, and \langle r^2 \rangle is the mean square radial distance of the electrons from the nucleus. This classical expression, derived from the Larmor theorem, applies to insulators and closed-shell systems where all orbitals are filled. Superconductors represent an extreme case of perfect diamagnetism, with susceptibility \chi = -1, expelling all magnetic flux from their interior. Paramagnetism occurs in materials with spins or orbital moments that align partially with an applied field, producing a positive . For localized moments, such as in ionic solids with ions like compounds, the follows at high temperatures: \chi = \frac{C}{T}, where C is the Curie , proportional to the square of the effective , and T is the . This law, experimentally established by in 1895, reflects the thermal randomization of spins against field alignment. In metals, an additional contribution comes from the Pauli paramagnetism of the free electron gas, where spin polarization at the enhances independently of . Ferromagnetism emerges from cooperative alignment of atomic moments below a critical temperature, leading to without an external field. The underlying mechanism is the , arising from the , which favors parallel spins to minimize spatial overlap of electrons on neighboring atoms. In the Heisenberg model, this is captured by the term -J \mathbf{S}_i \cdot \mathbf{S}_j, where J > 0 is the ferromagnetic exchange energy. Pierre Weiss introduced the in 1907, positing molecular fields that align moments within domains—regions of uniform magnetization—to explain bulk . The transition to occurs at the T_c, given in as k_B T_c = \frac{2 z J S(S+1)}{3}, with z the number of nearest neighbors and S the spin quantum number. Examples include iron (T_c \approx 1043 K) and nickel (T_c \approx 627 K), where exchange stabilizes long-range order. In ferromagnets, magnetization reversal exhibits hysteresis, where the magnetization lags behind the applied field due to domain wall motion and pinning. Domains form to minimize magnetostatic energy, with walls separating regions of opposite magnetization. The Barkhausen effect manifests as discrete jumps in magnetization during hysteresis, producing audible noise when amplified, as abrupt domain wall displacements occur under slowly varying fields. This irreversible behavior underpins applications in magnetic storage and transformers.

Dielectric Response and Optical Absorption

The dielectric response of solids to an applied electric field is characterized by the dielectric function \epsilon(\omega), which describes the material's polarization as a function of frequency \omega. This function is defined as \epsilon(\omega) = 1 + \chi(\omega), where \chi(\omega) is the electric susceptibility relating the induced polarization \mathbf{P} to the field via \mathbf{P} = \epsilon_0 \chi(\omega) \mathbf{E}. In solids, \epsilon(\omega) is complex, \epsilon(\omega) = \epsilon_1(\omega) + i \epsilon_2(\omega), with the imaginary part \epsilon_2(\omega) linked to absorption through \epsilon_2(\omega) = 4\pi \sigma(\omega)/\omega, where \sigma(\omega) is the conductivity. This formulation arises from Maxwell's equations and linear response theory, capturing how electrons and ions respond to electromagnetic waves. Polarization in dielectrics originates from several , each contributing to the \chi(\omega). Electronic polarization involves the displacement of electron clouds relative to atomic nuclei, dominant at high frequencies (optical range) and present in all insulators. Ionic polarization occurs in materials with ionic bonds, such as NaCl, where positive and negative ions shift relative to each other, contributing at frequencies. Orientational polarization arises in solids with permanent dipoles, like certain polymers or ferroelectrics, where dipoles align with the field, but it is typically slower and prominent at low frequencies ( range). These mechanisms combine to yield the total \mathbf{P} = N \mathbf{p}, where N is the of polarizable units and \mathbf{p} is the average . A key relation connecting macroscopic dielectric properties to microscopic polarizability is the Clausius-Mossotti equation, \frac{\epsilon - 1}{\epsilon + 2} = \frac{4\pi}{3} \frac{N \alpha}{V}, where \epsilon is the static constant, N/V is the of , and \alpha is the molecular . This equation derives from considering the local field acting on a molecule within a , assuming a spherical around it, and equates the macroscopic to the sum of induced dipoles. It holds for non-polar and relates to density via the Lorentz-Lorenz form for optical frequencies. Local field corrections refine the simple relation \mathbf{P} = N \alpha \mathbf{E} by accounting for the field from surrounding dipoles, leading to the Lorentz local field \mathbf{E}_{loc} = \mathbf{E} + \frac{4\pi}{3} \mathbf{P}. In solids, this correction is incorporated into the dielectric function as \epsilon(\omega) = \epsilon_{core} + \frac{4\pi i \sigma(\omega)}{\omega}, where \epsilon_{core} includes contributions from tightly bound electrons. These corrections are crucial in polar materials, enhancing splitting between transverse and longitudinal optical phonon modes, and are derived from Lorentz's cavity model excluding the molecule's own field. Optical absorption in solids arises primarily from interband transitions, where photons excite electrons across energy bands, particularly near the band gap E_g. For direct transitions in semiconductors like GaAs, vertical transitions in k-space conserve momentum, with the absorption rate governed by Fermi's golden rule: w_{abs} = \frac{2\pi}{\hbar} |\langle f | \hat{H}_{po} | i \rangle|^2 \delta(E_f - E_i - \hbar\omega). The joint density of states (JDOS), which counts accessible initial and final states, is \rho_{cv}(\hbar\omega) = \frac{1}{2\pi^2} \left( \frac{2\mu_{eff}}{\hbar^2} \right)^{3/2} (\hbar\omega - E_g)^{1/2} for parabolic bands near the minimum gap, leading to a characteristic square-root dependence in the absorption coefficient \alpha(\omega) \propto \sqrt{\omega - E_g / \hbar} just above the gap threshold. This form reflects the increasing number of available states as photon energy exceeds E_g. In insulators with large band gaps (E_g > 3 ), such as or fused silica, materials are transparent to since \hbar\omega (1.65–3.1 ) falls within the gap, resulting in negligible absorption and high (>90% over centimeters). Conversely, metals exhibit strong reflectivity (>90%) across the due to free carrier response, where intraband transitions cause rapid re-emission of incident , as described by the with negative real \epsilon_1(\omega) and large imaginary part. These properties distinguish insulators' optical clarity from metals' mirror-like reflection.

Modern Developments

Superconductivity Basics

Superconductivity manifests as the complete dissipationless flow of in certain materials below a critical T_c, where () electrical drops to zero. This phenomenon was first observed in 1911 by while studying the electrical resistivity of mercury cooled with , revealing a sharp transition to zero at approximately 4.2 K. A defining feature beyond zero is the , in which a superconductor expels nearly all magnetic fields from its interior upon entering the superconducting state, behaving as a perfect diamagnet. This expulsion was experimentally demonstrated in 1933 by and Robert Ochsenfeld using lead and tin samples, showing that the inside the material vanishes below T_c regardless of prior field exposure. Superconductors are categorized into type I and type II based on their behavior, a distinction arising from the Ginzburg-Landau \kappa = \lambda / \xi, where \lambda is the and \xi is the . Type I superconductors, typically pure metals like aluminum and lead with \kappa < 1/\sqrt{2}, exhibit complete field expulsion up to a critical field H_c beyond which they revert to the normal state. In contrast, type II superconductors, such as alloys like NbTi with \kappa > 1/\sqrt{2}, allow to penetrate in quantized vortices between a lower critical field H_{c1} and an upper critical field H_{c2}, enabling practical applications in high-field magnets. This mixed-state behavior was theoretically predicted by Alexei Abrikosov in 1957 as a of vortices that stabilizes the superconducting . The microscopic understanding of conventional superconductivity emerged from the Bardeen-Cooper-Schrieffer ( in 1957, which posits that electrons near the form loosely bound pairs, known as Cooper pairs, due to an attractive interaction mediated by phonons. These pairs condense into a coherent with a that overcomes the repulsion, resulting in a superconducting energy gap $2\Delta(0) for single-particle excitations, empirically related by $2\Delta(0) \approx 3.5 k_B T_c at zero temperature. The phonon-mediated nature was confirmed by the isotope effect, where the critical temperature scales inversely with the ionic mass as T_c \propto M^{-1/2}, first measured in mercury isotopes in 1950, indicating that vibrations are essential for pairing. Phenomenologically, the electromagnetic properties are captured by the London equations, introduced by Fritz and Heinz London in 1935 to explain the Meissner effect through a two-fluid model. The first equation relates the supercurrent density to the vector potential as \mathbf{J} = -\frac{n_s e^2}{m} \mathbf{A}, where n_s is the density of superconducting electrons, implying perfect screening; combined with Maxwell's equations, it yields exponential decay of magnetic fields with penetration depth \lambda = \sqrt{\frac{m}{\mu_0 n_s e^2}}, typically on the order of 10-100 nm in conventional superconductors. A major advance came in 1986 with the discovery of high-T_c superconductivity in copper oxide (cuprate) materials by J. Georg Bednorz and K. Alex Müller, achieving T_c > 30 K in the La-Ba-Cu-O system, far exceeding the BCS limit for phonons and involving unconventional pairing mechanisms beyond simple electron-phonon coupling. As of 2025, advances include high-entropy superconductors and stabilization of high-Tc phases at ambient pressure.

Topological Materials and Nanostructures

Topological materials represent a class of quantum where the electronic band structure exhibits nontrivial topological properties, leading to robust surface or edge states protected by symmetry. These properties arise from global invariants in the bulk band structure, distinguishing them from conventional materials where transport is governed by local disorder-sensitive features. In , topological materials have revolutionized understanding of , enabling dissipationless edge transport and spin-polarized currents, with applications in and . Topological insulators are a paradigmatic example, characterized by an insulating but conducting due to the nontrivial of the bulk bands. The topological nature is quantified by the \mathbb{Z}_2 invariant, which distinguishes trivial insulators (invariant 0) from nontrivial ones (invariant 1), preserved under time-reversal symmetry. These surface states form a single , with spin-momentum locking where the is perpendicular to its , suppressing backscattering and enabling helical . Bismuth selenide (Bi_2Se_3) serves as a prototypical three-dimensional topological insulator, predicted theoretically in 2009 and experimentally confirmed via , exhibiting a of approximately 0.3 eV with protected . The Chern number, originally from quantum Hall physics, relates to the curvature integral over the but in time-reversal-invariant systems like topological insulators, it pairs to yield the \mathbb{Z}_2 index. The provides foundational insight into topological phases, observed in two-dimensional gases under strong perpendicular magnetic fields. In the integer quantum Hall effect (IQHE), discovered in 1980, the Hall conductivity quantizes as \sigma_{xy} = \frac{ne^2}{h}, where n is an integer, e the charge, and h Planck's constant, arising from filled forming chiral edge states topologically protected against impurities. The (FQHE), observed in 1982 at filling factors like \nu = 1/3, extends this to strongly interacting s, where quasiparticles with fractional charge e/3 emerge in an incompressible fluid state, explained by Laughlin's wavefunction. These effects underscore how enforces quantized , independent of sample details. Nanostructures in solid-state physics exploit quantum confinement to engineer topological features in reduced dimensions. In zero-dimensional quantum dots, electrons are confined in all directions, leading to discrete energy levels where the confinement energy scales as E \propto 1/L^2, with L the characteristic size, altering optical and electrical properties for applications like quantum bits. One-dimensional quantum wires and two-dimensional quantum wells similarly quantize motion, enhancing density of states and enabling tunable band gaps. Graphene, a two-dimensional nanostructure, exemplifies this with its honeycomb lattice yielding massless Dirac fermions at the K and K' points, where the dispersion is linear E = \hbar v_F |k|, v_F the Fermi velocity, resulting in exceptional electron mobility over 200,000 cm²/Vs. Post-2020 advances have expanded topological materials into novel regimes, including Weyl semimetals and engineered two-dimensional systems. Weyl semimetals feature bulk Weyl nodes—monopole-like sources of Berry curvature—acting as three-dimensional analogs of Weyl fermions, with TaAs as an early realization exhibiting Fermi arc surface states. Recent progress includes low-symmetry Weyl semimetals like , where tilted cones enable type-II Weyl points, enhancing effects for thermoelectric applications. In twisted , rotating layers by the of approximately 1.1° flattens bands near the Dirac points, fostering strongly correlated states like unconventional observed since 2018, with ongoing 2023–2025 studies revealing topological Chern insulators tunable by strain and doping. As of 2025, high-throughput searches have identified new magnetic topological materials, and studies on twisted have revealed exotic quantum phenomena. These developments highlight the integration of with interactions, addressing gaps in earlier models by incorporating moiré superlattices for designer .

References

  1. [1]
    Introduction to Solid State Physics
    What is Solid State Physics? Crystalline Solids: Symmetry and Bonding · Experimental Determination of Crystal Structures · Electronic Band Structure of Solids ...
  2. [2]
    Introduction to Solid-State Physics - UCLA General Catalog
    Introduction to basic theoretical concepts of solid-state physics with applications. Crystal symmetry; cohesive energy; diffraction of electron, neutron, and ...
  3. [3]
    The Division of Condensed Matter Physics
    Condensed matter physics studies matter's properties, from atoms to tens of centimeters, involving electrons, ions, and photons, and includes both solids and ...
  4. [4]
    Physics 140A: Introduction to solid state physics - Vishik Lab
    Physics 140A is an introductory course to the physics of crystalline solids, to prepare students for understanding applications of this field.
  5. [5]
    Physics 406: INTRO TO SOLID STATE PHYSICS
    Overview: This course is intended to provide an introduction to the physics of solids. We will begin by characterizing the properties of static (crystal ...<|control11|><|separator|>
  6. [6]
  7. [7]
    Some basic concepts in solid state physics - Book chapter
    This chapter introduces some of the most important concepts of solid state physics, due to Drude and Sommerfeld, to explain the electrical and optical ...
  8. [8]
    Introduction to Solid State Physics, 8th Edition
    ### Summary of *Introduction to Solid State Physics, 8th Edition*
  9. [9]
    APS -APS March Meeting 2020 - Event - Solid State Physics across ...
    Solid state physics encompasses a variety of physics phenomena, including modern ones such as high temperature superconductivity and nanotechnology.
  10. [10]
    Solid State Physics - Principedia - Princeton University
    Solid-state physics is the study of the properties of solid materials and how they emerge quantum mechanically from their constituent atoms. It is the ...
  11. [11]
    Solid State Physics
    Solid state physics is the study of such emergent phenomena. The primary question is: "How does chemical structure and quantum mechanics come together?
  12. [12]
    [PDF] PHYS 666: Solid State Physics I
    Condensed-matter physics is the more modern term. • Condensed-matter physics is broader and applies to concepts that work in solids, but could.
  13. [13]
    [PDF] How Chemistry and Physics Meet in the Solid State - FZU
    To make sense of the marvelous electronic properties of the solid state, chemists must learn the language of solid-state physics, of band structures.
  14. [14]
    Categories of Solids
    Amorphous solids (literally, "solids without form") have a random structure, with little if any long-range order. Polycrystalline solids are an aggregate of a ...
  15. [15]
    [PDF] The Crystalline Solid State
    Oct 19, 2015 · Crystalline solids have atoms/ions/molecules arranged in regular, repeating patterns. They possess long-range periodicity.
  16. [16]
    Quantum Numbers and Electron Configurations
    Quantum Numbers. The Bohr model was a one-dimensional model that used one quantum number to describe the distribution of electrons in the atom.
  17. [17]
    The reflection of X-rays by crystals - Journals
    The reflection of X-rays by crystals. William Henry Bragg, Proceedings A, 1913 · The analysis of crystals by the X-ray spectrometer. William Lawrence Bragg, ...
  18. [18]
    1947: Invention of the Point-Contact Transistor | The Silicon Engine
    John Bardeen & Walter Brattain achieve transistor action in a germanium point-contact device in December 1947.
  19. [19]
    DCMP and DMP Celebrate 50 Years - American Physical Society
    ... Physics was formed in 1944.) In 1978 the DSSP was renamed the Division of Condensed Matter Physics to recognize that disciplines covered in the division ...
  20. [20]
  21. [21]
    When condensed-matter physics became king
    Jan 1, 2019 · In 1978, APS's division of solid state physics became the division of condensed matter physics. The new name offered self-identified ...
  22. [22]
    Gartner Says Worldwide Semiconductor Revenue Declined 11% in ...
    Jan 16, 2024 · Worldwide semiconductor revenue in 2023 totaled $533 billion, a decrease of 11.1% from 2022, according to preliminary results by Gartner, Inc.
  23. [23]
    The rise of quantum materials | Nature Physics
    Feb 2, 2016 · Their study now extends beyond strongly correlated electron systems, giving rise to the broader concept of quantum materials. Measured in terms ...
  24. [24]
    The 4 Reciprocal Lattice - International Union of Crystallography
    This concept and the relation of the direct and reciprocal lattices through the Fourier transform was first introduced in crystallography by P. P. Ewald (. ...
  25. [25]
    Origin of the stabilized simple-cubic structure in polonium: Spin-orbit ...
    Among the elements in the periodic table, only Po (polonium) is known to crystallize in the simple-cubic (SC) structure in nature. 1. Radioactive Po, which was ...
  26. [26]
    General and Atomic Properties of Copper
    Crystallographic Features of Copper ; Crystal Structure, face-centered cubic ; Number of Atoms per Unit Cell, 4 ; Lattice Parameters at 293 K, 3.6147 x 10 -10 m.Missing: confirmation | Show results with:confirmation
  27. [27]
    mp-13: Fe (Cubic, Im-3m, 229) - Materials Project
    Fe is Tungsten structured and crystallizes in the cubic Im̅3m space group. Fe is bonded in a distorted body-centered cubic geometry to eight equivalent Fe atoms.
  28. [28]
    Crystallography - Group Theory in Physics
    Looking at the rotations that are elements of these point groups, one can conclude that there are only 32 crystallographic point groups in three dimensions.
  29. [29]
    [PDF] Crystallographic Point Groups and Space Groups
    Jun 8, 2011 · ▷ 32 distinct point groups (10 in 2D). ▷ 230 distinct space groups (17 ... Space Groups for Solid State Scientists. Academic Press, 1990.
  30. [30]
    [PDF] CHAPTER 4: IMPERFECTIONS IN SOLIDS
    Defects affect material properties (e.g., grain boundaries control crystal slip). • Defects may be desirable or undesirable. (e.g., dislocations may be good or ...Missing: physics | Show results with:physics
  31. [31]
    Schottky Defects - Chemistry LibreTexts
    Jun 30, 2023 · A Schottky defect occurs when oppositely charged ions vacate their sites, and for MX compounds, one cation and one anion leave their sites.
  32. [32]
    Further calculations of the energies of formation of Schottky defects ...
    The author presents the results of calculations of the energy of formation of Schottky defects in some alkali halides and alkaline earth oxides. It is shown ...Missing: concentrations | Show results with:concentrations
  33. [33]
    Crystal Defect - an overview | ScienceDirect Topics
    The 0-dimensional are point defects which include vacancies, interstitial atoms, substitutional atoms, and antisites (in compounds). If substitutional atoms are ...
  34. [34]
    [PDF] chapter 5: diffusion in solids
    To make solid-diffusion practical, point defects (e.g. vacancies) are generally required. • Or/and migrating atoms that are small enough to fit into the ...
  35. [35]
    Interfacial plasticity governs strain rate sensitivity and ductility in ...
    Interfacial plasticity governs strain rate sensitivity and ductility in nanostructured metals ... Contributions of bulk dislocation are required in the absence of ...
  36. [36]
    X-ray diffraction, Bragg's law and Laue equation
    Sep 7, 2021 · Very strong intensities known as Bragg peaks are obtained in the diffraction pattern when scattered waves satisfy the Bragg's Law. Following ...
  37. [37]
    [PDF] Chapter 3 X-ray diffraction • Bragg's law • Laue's condition
    An incident wave vector will satisfy the Laue condition if and only if the tip of the vector lies in a plane that is the perpendicular bisector of a line ...
  38. [38]
    [PDF] X-ray Diffraction (XRD)
    2θ-diffraction angle. Diffraction occurs only when Bragg's Law is satisfied Condition for constructive interference (X-rays 1 & 2) from planes with spacing d.
  39. [39]
    Electron-phonon interactions from first principles | Rev. Mod. Phys.
    Feb 16, 2017 · This article reviews the theory of electron-phonon interactions in solids from the point of view of ab initio calculations.
  40. [40]
  41. [41]
    [PDF] Relation between compressibility, thermal expansion, atom volume ...
    Subsequent to the compressibility measurements (Grüneisen 1907a+b, 1908a+b) I therefore tested the theory on metals who are occasionally considered monatomic.
  42. [42]
    Zur kinetischen Theorie der Wärmeleitung in Kristallen - Peierls - 1929
    Zur kinetischen Theorie der Wärmeleitung in Kristallen. R. Peierls,. R ... Download PDF. back. Additional links. About Wiley Online Library. Privacy Policy ...
  43. [43]
    Zur Elektronentheorie der Metalle auf Grund der Fermischen Statistik
    Cite this article. Sommerfeld, A. Zur Elektronentheorie der Metalle auf Grund der Fermischen Statistik. Z. Physik 47, 1–32 (1928). https://doi.org/10.1007 ...
  44. [44]
    Über Gasentartung und Paramagnetismus - SpringerLink
    Pauli Jr. Zeitschrift für Physik A Hadrons and nuclei volume 41, pages 81–102 (1927)Cite this article ... June 1927. DOI : https://doi.org/10.1007/BF01391920 ...
  45. [45]
    [PDF] Über die Quantenmechanik der Elektronen in Kristallgittern
    Von Felix Bloeh in Leipzig. 5lit 2 Abbildungen. (Eingegangen am 10. August 1928.) Die Bewegung eines Elektrons im Gitter wird untersucht, indem wir uns dieses.
  46. [46]
    [PDF] Léon Brillouin and the Brillouin Zone - Physics Courses
    Abstract. In this report, I followed Brillouin's derivation of energy gap near degeneracy and compared it with contemporary formulation.
  47. [47]
    2.2: Bands of Orbitals in Solids - Chemistry LibreTexts
    Apr 8, 2021 · The bands of orbitals arising in any solid lattice provide the orbitals that are available to be occupied by the number of electrons in the crystal.
  48. [48]
  49. [49]
    Band Gap – dielectrics, semiconductors, metals, energy, electronic ...
    Examples of direct band gap semiconductor materials are gallium arsenide (GaAs), indium gallium arsenide (InGaAs), gallium nitride (GaN), aluminum nitride (AlN) ...What are Electronic Band Gaps? · Semiconductors · Methods for Measuring Band...
  50. [50]
    Direct and Indirect Band Gap Semiconductors
    In a direct band gap semiconductor, the top of the valence band and the bottom of the conduction band occur at the same value of momentum.
  51. [51]
    2.4: Band Theory of Solids - Workforce LibreTexts
    Mar 19, 2021 · Since the valence and conduction bands overlap in metals, little energy removes an electron. Metals are excellent conductors. The large gap ...
  52. [52]
    NSM Archive - Diamond (C) - Band structure and carrier concentration
    The values as low as mν=0.16mo (Kemmey and Wederpohl [1965]) and as high as mν=1.1mo (Dean [1965]) have been reported. For estimations, one can use the value ...Missing: source | Show results with:source
  53. [53]
    4. Fermi Energy Levels - Engineering LibreTexts
    Jul 5, 2021 · On the other hand, insulators and semiconductors have Fermi levels lying in the forbidden band gap and have full valence bands, therefore ...
  54. [54]
    Band Theory and Fermi Level - EdTech Books
    The position of the Fermi level in a semiconductor is temperature-dependent. At absolute zero, the Fermi level is fixed and remains midway between the ...
  55. [55]
    Magic of high-order van Hove singularity | Nature Communications
    Dec 18, 2019 · The van Hove singularity in density of states generally exists in periodic systems due to the presence of saddle points of energy dispersion ...
  56. [56]
    The discovery of three-dimensional Van Hove singularity - Nature
    Mar 14, 2024 · Van Hove singularity (VHS) originates from the critical point in the electronic band structure with divergence in the density of states (DOS).
  57. [57]
    [PDF] Intrinsic carrier concentration in semiconductors - Galileo
    Intrinsic carrier concentration in semiconductors. Melissinos, eq.(1.4), gives the formula, valid at thermal equilibrium, ni = Ns exp(¡. Eg. 2kBT \. (1) where ...
  58. [58]
    [PDF] Lecture 4: Intrinsic semiconductors - An-Najah Staff
    carrier concentration can be calculated. ni = p. NcNv exp(−. Eg. kBT. ) (12). Thus, the intrinsic carrier concentration is a semiconductor is dependent only on ...
  59. [59]
    [PDF] 1. Carrier Concentration a) Intrinsic Semiconductors
    In an intrinsic semiconductor, the intrinsic carrier concentration (ni) is the concentration of electrons in the conduction band, equal to the concentration of ...
  60. [60]
    [PDF] 19700026421.pdf - NASA Technical Reports Server (NTRS)
    Jun 1, 1970 · Phwphbrus has a donor level in silicon approximately 0.044 eV below the conduction band edge. This donor level is actually a multiplet, as ...
  61. [61]
    [PDF] NASA TN 0-7517
    dp/dT _ -dpp/dT. (4). P. Mp. The ionization energy of boron in silicon is 0.045 eV, and most of the boron is ionized at temperatures well below room temperature ...
  62. [62]
    [PDF] Conduction in Semiconductors
    (1.8). This relation is called the mass-action law. To understand this equation, consider an intrinsic semiconductor in which n = p = ni. Assume.
  63. [63]
    [PDF] Lecture-2 - Electrical Engineering
    For a pure semiconductor, n = p = ni is the intrinsic concentration, the ... of the semiconductor is proportional to the concentration of free carriers, σ may.Missing: formula | Show results with:formula<|control11|><|separator|>
  64. [64]
    [PDF] Extrinsic Semiconductors - MIT OpenCourseWare
    Expected Temperature Behavior of Doped. Material (Example:n-type). • 3 regimes. E g. /2k b ln(n). E b. /k b. Intrinsic Extrinsic. Freeze-out. 1/T. ©1999 E.A. ...
  65. [65]
    [PDF] Electrons and Holes in Semiconductors
    Semiconductors doped with acceptors have many holes and few mobile electrons, and are called P type because holes carry positive (P) charge. Boron is the most.
  66. [66]
    [PDF] Drude, P., 1900, Annalen der Physik 1, 566
    Zur Elektronentheorie der Metalle; von P. Drude. I. Teil. Dass die Elektricitätsleitung der Metalle ihrem Wesen nach nicht allzu verschieden von der der ...
  67. [67]
    Zur Elektronentheorie der Metalle; II. Teil. Galvanomagnetische und ...
    Annalen der Physik · Volume 308, Issue 11 pp. 369-402 Annalen der Physik ... Drude, Physik. Zeitschr. 1. p. 161. 1900. Google Scholar. Citing Literature.
  68. [68]
    Hall Effect - Engineering LibreTexts
    Sep 7, 2021 · Hall Effect, deflection of conduction carriers by an external magnetic field, was discovered in 1879 by Edwin Hall.
  69. [69]
    [PDF] The Hall Effect
    Define the hall coefficient as RH ≡ Ey/(BJx) and show that RH = 1/(nq). Note that this model predicts two things: the Hall coefficient is independent of the ...Missing: nec) | Show results with:nec)
  70. [70]
    [PDF] The Hall Effect
    A convenient experimental quantity is the Hall coefficient, RH defined as: RH = Ey. JxB. = eBτEx m m. ne2τExB. = 1 ne. (13). The Hall coefficient is positive if ...
  71. [71]
    [PDF] 2.5 THE HALL EFFECT AND HALL DEVICES - PhysLab
    (1) J₂ B₂. JxBz. = Jx B₂/n or. [2.30]. A useful parameter called the Hall coefficient RH is defined as. RH = Ey. JxBz. [2.31]. The quantity RH measures the ...
  72. [72]
    [PDF] Hall Effect Measurement Handbook - Quantum Design
    A measurement of the. Hall effect and the resistivity provides a wealth of information such as the carrier density, the mobility of the carriers, and the ...Missing: coolers | Show results with:coolers<|separator|>
  73. [73]
    [PDF] Thermoelectric Energy Conversion - DSpace@MIT
    The two most important thermoelectric effects are the Seebeck effect and the Peltier effect. 1.1.1 Seebeck Effect and Phonon drag. As shown in Figure 1-1a ...
  74. [74]
    Synergetic Enhancement of Seebeck Coefficients and Electrical ...
    Mar 1, 2023 · ... Peltier effect. The figure of merit (ZT, dimensionless quantity), ZT = S2σT/k, of TE materials can be used to evaluate their energy ...
  75. [75]
    Relationship between thermoelectric figure of merit and energy ...
    The dimensionless figure of merit, ZT = S2ρ−1κ−1T, is calculated from the Seebeck coefficient (S), electrical resistivity (ρ), and thermal conductivity (κ).Missing: σ | Show results with:σ
  76. [76]
    [PDF] Boltzmann Transport - Physics Courses
    Boltzmann equation in the relaxation time approximation has the solution δf = −τ(ε)v ·. eS + ε − µ. T. VT. −. ∂f0. ∂ε . (1.178). We now consider both the ...
  77. [77]
    [PDF] Lecture 32: Introduction to Boltzmann Transport - MIT
    Apr 28, 2004 · • Boltzmann Transport Equation. • Relaxation Time Approximation Overview. • Example: Low-field Transport in a Resistor. Outline. April 28,2004.
  78. [78]
    Developing a two-parabolic band model for thermoelectric transport ...
    Aug 4, 2022 · In this study, we present a systematic methodology for developing a two-band model using one valence and one conduction band for any given TE material.Missing: sensors | Show results with:sensors
  79. [79]
    Polycrystalline Parametrized as a Narrow-Band-Gap Semiconductor ...
    Jan 24, 2018 · Using the two-band model, the thermoelectric performance at different doping levels is predicted, finding and 0.1 for and type, respectively ...Missing: sensors | Show results with:sensors
  80. [80]
    [PDF] SOLID STATE PHYSICS PART III Magnetic Properties of Solids - MIT
    All atoms in gases, liquids and solids exhibit a diamagnetic contribution to the total sus- ... The diamagnetic susceptibility is then [χ = (∂~µ/∂ ~H)/volume] χ = ...
  81. [81]
    [PDF] A rigorous derivation of the Larmor and Van Vleck contributions. - HAL
    Apr 24, 2014 · This led to the so-called Langevin formula for the diamagnetic susceptibility per unit volume of n electrons, see [38, pp. 89]: χdia. La ...
  82. [82]
    Meissner Effect - Engineering LibreTexts
    Sep 7, 2021 · The magnetic susceptibility equals -1, meaning the superconductor has perfect diamagnetism. A Type I superconductor has the Meissner state ...
  83. [83]
    [PDF] The Specific Heat of Solids, Curie Paramagnetism
    Feb 13, 2021 · ... Law of Paramagnetism. (2.13.27). This χ ∼ 1/T is known as the Curie law of paramagnetism, discovered experimentally by Pierre Curie in ∼ 1895.
  84. [84]
    A Survey of the Theory of Ferromagnetism | Rev. Mod. Phys.
    P. Weiss, J. de Phys. 6, 667 (1907); P. Langevin, J. de Phys. 4, 678 (1905) Chap. II of the writer's Electric and Magnetic Susceptibilities (Oxford ...Missing: paper | Show results with:paper
  85. [85]
    [PDF] Mean field theory - bingweb
    Apr 18, 2018 · The Curie-Weiss law, the temperature dependence of the spontaneous magnetization, and the mean-field exponents of the magnetization and magnetic ...Missing: domains | Show results with:domains
  86. [86]
    The magnetic barkhausen effect - Taylor & Francis Online
    The effect was named after its discoverer who, in 1919, connected a coil surrounding a ferromagnet to a newly invented amplifier and listened to the output of ...
  87. [87]
    [PDF] SOLID STATE PHYSICS PART II Optical Properties of Solids - MIT
    The dielectric function ε(ω) approaches the static dielectric constant ε0 as ω → 0. Also, ε(ω) approaches the high frequency dielectric function ε∞ as ω ...
  88. [88]
    Polarisation mechanisms - DoITPoMS
    There are three main polarisation mechanisms that can occur within a dielectric material: electronic polarisation, ionic polarisation (sometimes referred to as ...
  89. [89]
    [PDF] On the Clausius-Mossotti Relation - Kirk T. McDonald
    May 20, 2025 · In this note we review the derivation of the Clausius-Mossotti relation [1, 3] between the molecular polarizability α and the relative ...
  90. [90]
    [PDF] 32 Optical absorption in semiconductors - David Miller
    Joint density of states. We can then define a “joint density of states” the number of transitions per unit (photon) energy where we view E. J as being a ...
  91. [91]
    Transparency – light, absorption, scattering, dielectric materials ...
    Metals reflect light. Transparency for visible light is not possible for metals (except for extremely thin layers), mostly because light is strongly reflected.
  92. [92]
    The super century | Nature Materials
    Mar 24, 2011 · On 8 April 1911, Heike Kamerlingh Onnes and his assistants were measuring the low-temperature electrical properties of mercury. As soon as ...
  93. [93]
    [PDF] Meissner and Ochsenfeld revisited - Physics Courses
    The dis- covery described in that paper represents a major and crucial landmark in the development of our understanding of superconductivity and it would seem ...
  94. [94]
    Theory of Superconductivity | Phys. Rev.
    A theory of superconductivity is presented, based on the fact that the interaction between electrons resulting from virtual exchange of phonons is attractive.
  95. [95]
    Superconductivity of Isotopes of Mercury | Phys. Rev.
    This paper, 'Superconductivity of Isotopes of Mercury', by C. A. Reynolds, B. Serin, W. H. Wright, and L. B. Nesbitt, was published in Phys. Rev. 78, 487 on ...Missing: effect | Show results with:effect
  96. [96]
    [PDF] The electromagnetic equations of the supraconductor
    This equation, which might replace Ohm's law for supraconductors, simply expresses the influence of the electric part of the Lorentz force on freely movable ...Missing: brothers superconductivity URL
  97. [97]
    Topological Insulators in Three Dimensions | Phys. Rev. Lett.
    Mar 7, 2007 · Topological Insulators in Three Dimensions. Liang Fu, C. L. Kane, and E. J. Mele. Department of Physics and Astronomy, University of ...
  98. [98]
    Topological insulators with inversion symmetry | Phys. Rev. B
    In this paper, we show that the presence of inversion symmetry greatly simplifies the problem of evaluating the invariants. We show that the invariants can be ...Abstract · Article Text · INVARIANTS IN TWO AND... · TOPOLOGICAL PHASES IN...
  99. [99]
    [PDF] THE QUANTIZED HALL EFFECT - Nobel lecture, December 9, 1985
    THE QUANTIZED HALL EFFECT. Nobel lecture, December 9, 1985 by. KLAUS von KLITZING,. Max-Planck-Institut für Festkörperforschung, D-7000 Stuttgart 80. 1 ...