Fact-checked by Grok 2 weeks ago

Potential well

A potential well is a in a potential energy landscape where the potential energy V(x) for a particle is lower than in the adjacent regions, confining the particle's motion if its total energy E is less than the surrounding barrier height, as visualized in graphs of V(x) versus position x. This concept applies across classical and , with the well's shape—such as square, harmonic, or more complex forms—determining the allowed behaviors and energy states of the particle. In , a particle in a potential well oscillates between turning points where its E - V(x) = 0, remaining confined within the classically allowed region where E > V(x) and unable to enter the forbidden regions where E < V(x). Quantum mechanically, the time-independent Schrödinger equation -\frac{\hbar^2}{2m} \frac{d^2 \psi}{dx^2} + V(x) \psi = E \psi governs the system, yielding discrete bound states with quantized energies for E below the barrier, and wavefunctions that decay exponentially in forbidden regions due to tunneling effects. For the idealized infinite square well, where V(x) = 0 inside a finite interval and V(x) = \infty outside, the energy levels are E_n = \frac{n^2 \pi^2 \hbar^2}{2 m a^2} (with n = 1, 2, 3, \dots and well width a), and the wavefunction vanishes at the boundaries. In contrast, a finite square well (with V(x) = 0 inside and V(x) = V_0 > 0 outside) allows a finite number of bound states depending on well depth V_0 and width, solved via transcendental equations like \tan(k L/2) = \gamma / k for even solutions, where k = \sqrt{2 m E}/\hbar and \gamma = \sqrt{2 m (V_0 - E)}/\hbar. Potential wells model fundamental physical systems, including orbitals in atoms and molecules, where attractive potentials form effective wells binding particles. In , the strong force creates a finite well approximating confinement within the , with well matching nuclear size and depths around 50 MeV. Double-well potentials describe phenomena like inversion via quantum tunneling between symmetric minima. In modern applications, engineered quantum wells in semiconductor heterostructures (e.g., GaAs/AlGaAs layers) confine s and holes to nanometer scales, enabling devices like low-threshold lasers, optical modulators, and LEDs by exploiting quantized energy levels and reduced . These structures highlight quantum confinement effects, altering material properties for and .

Fundamentals

Definition and Characteristics

A potential well is a in space where the of a particle is lower than in the surrounding areas, effectively creating a trap that confines particles whose total energy is less than the height of the enclosing potential barriers. This configuration arises in various physical systems, where the function exhibits a local minimum, preventing particles with insufficient from escaping the . Key characteristics of a potential well include its depth, defined as the difference between the minimum at the well's bottom and the height of the surrounding barriers; its width, which measures the spatial extent of the low-potential region; its shape, such as parabolic for oscillators or rectangular for idealized models; and its dimensionality, ranging from one-dimensional wells along a line to three-dimensional wells in full space. These properties determine the well's ability to sustain bound states, where particles oscillate or remain confined within without escaping. Physically, potential wells lead to the confinement of particles lacking the energy to surmount the barriers, resulting in stable bound states essential for phenomena like planetary in gravitational wells or binding in electrostatic wells. For instance, Earth's forms a potential well that traps satellites in if their energy is below the . The concept of the potential well originates from 19th-century classical physics, rooted in potential theory developed by mathematicians such as and , who formalized the mathematical description of gravitational and electrostatic potentials through equations linking potential to mass or charge distributions. This framework laid the groundwork for understanding how potential landscapes govern particle motion in conservative force fields.

Visual and Conceptual Analogies

Potential wells are commonly visualized through two-dimensional plots where the V(x) is graphed against x, depicting the well as a "" in an landscape with the particle's represented as a horizontal line above the minimum. For a potential well, the curve appears as a smooth, U-shaped parabola, illustrating symmetric confinement, while a square well is shown as a rectangular dip with vertical walls, emphasizing abrupt boundaries. These diagrams highlight how particles oscillate within the well between turning points where is zero, providing an intuitive grasp of bounded motion. A classic conceptual for a potential well is a rolling in a , where the bowl's curved surface represents the profile, and the marble's position corresponds to the particle's location, demonstrating classical trapping and oscillatory behavior as the marble seeks the lowest point. Similarly, a in a illustrates how confines the object, with escape requiring sufficient to surmount the surrounding hills, mirroring the conditions for particle liberation from the well. In a quantum context, an bound in an can be likened to a particle in a potential well, where the "holding" effect arises from electrostatic forces rather than mechanical constraints, underscoring binding without classical contact. Visualization tools enhance understanding of potential wells, including energy diagrams that overlay particle trajectories on V(x) plots and contour plots for multi-dimensional wells, such as radial potentials V(r) in central force problems, which reveal circular or elliptical basins. Software like facilitates plotting these profiles, allowing users to generate custom curves for various well shapes and simulate particle dynamics interactively. A common misconception is viewing potential wells as literal physical objects or holes in space, rather than abstract representations of landscapes shaped by forces; this linguistic from the "well" can lead to confusion with tangible containers. Another error involves conflating potential wells with force fields, overlooking that the well describes variation, while forces derive from its , potentially misguiding intuitions about particle interactions.

Mathematical Description

Potential Energy Profile

The potential energy profile of a potential well is graphically represented by plotting the V(x) as a of x in one or V(r) as a of radial r in higher dimensions, revealing the spatial variation that governs particle confinement. Key features include minima, which correspond to stable equilibrium points where F = -\frac{dV}{dx} = 0 and the second \frac{d^2V}{dx^2} > 0, indicating restorative behavior; maxima, representing energy barriers that particles must overcome to escape; and points, where \frac{d^2V}{dx^2} = 0, often marking transitions to unstable regions. These profiles are typically sketched with the horizontal axis as and the vertical axis as , showing bounded regions below a certain where classical particles are trapped. Common functional forms for potential wells include the approximation V(x) = \frac{1}{2} k x^2, which describes small oscillations around with a parabolic symmetric about the minimum at x = 0, widely used for vibrational modes in molecules and solids. The potential, V(x) = -\frac{k}{|x|} in one dimension or V(r) = -\frac{k}{r} radially, forms an asymmetric well diverging to -\infty as x \to 0 and approaching zero from below at large distances, modeling electrostatic interactions in atomic systems. For more realistic molecular bonds, the V(r) = D_e \left(1 - e^{-a(r - r_e)}\right)^2 captures with a finite depth, steep rise near equilibrium separation r_e, and gradual flattening at dissociation, better approximating anharmonic effects than the harmonic form. Characteristic parameters of a potential well include the well depth \Delta V, defined as the energy difference from the minimum to the asymptotic or barrier height, quantifying strength; the width L, often the full width at half-depth or between inflection points, influencing the number of bound states; and the at the minimum \frac{d^2V}{dx^2} \big|_{x=0} = k, which relates to the oscillation frequency \omega = \sqrt{k/m} for small amplitudes in . These parameters scale the profile's overall shape and depth, with the number of bound states in generally increasing for deeper and wider wells. In multi-dimensional cases, particularly for central force problems, the potential extends to radial form V(r) in spherical coordinates, where the effective potential becomes U(r) = V(r) + \frac{\ell(\ell+1)\hbar^2}{2m r^2} incorporating centrifugal effects, with minima shifted outward for nonzero \ell. This radial profile facilitates in the , enabling analysis of orbital motion in systems like atomic or gravitational wells.

Governing Equations

In , the behavior of a particle in a potential well is governed by Newton's second law, which relates the to the of the function V(x). The acting on a particle of m is F = -\frac{dV}{dx}, leading to the equation of motion m \frac{d^2 x}{dt^2} = -\frac{dV}{dx}. This describes the of the particle under the conservative derived from the potential, assuming no dissipative effects. For conservative systems, the dynamics can be reformulated using the H = \frac{p^2}{2m} + V(x), where p is the conjugate to x. This total energy function provides a phase-space description, with the first term representing and the second . Energy conservation follows directly, as the total energy E = \frac{1}{2} m v^2 + V(x) remains constant along the trajectory, where v = \frac{dx}{dt}. In a potential well, for cases with finite surrounding barriers, bound motion occurs when E is less than the barrier height, confining the particle to oscillatory motion within the well; in potentials where V \to +\infty as |x| \to \infty, such as the , the particle is bound for any finite E. The Hamiltonian formulation yields Hamilton's equations for phase-space evolution: \dot{x} = \frac{\partial H}{\partial p} and \dot{p} = -\frac{\partial H}{\partial x}. These canonical equations, symmetric in form, facilitate analysis of trajectories and conserved quantities in potential wells, such as periodic orbits for bound motion. Transitioning to the quantum regime, the governing equation becomes the time-independent Schrödinger equation \hat{H} \psi = E \psi, where the Hamiltonian operator is \hat{H} = -\frac{\hbar^2}{2m} \frac{d^2}{dx^2} + V(x). Here, \psi(x) is the wavefunction, \hbar is the reduced Planck's constant, and m is the particle mass, with solutions normalized such that \int |\psi|^2 dx = 1 to ensure probabilistic interpretation. The kinetic term involves \hbar, setting the scale for quantum effects, while m influences the curvature of the wavefunction in regions of varying V(x).

Classical Perspective

Dynamics of Trapped Particles

In , particles trapped within a potential well exhibit bound motion when their total E is less than the of the surrounding potential barrier. The particle oscillates periodically between two turning points, where its vanishes and the equals E. These turning points, denoted x_1 and x_2, satisfy V(x_1) = V(x_2) = E, with the motion confined to the interval x_1 < x < x_2. The period T of this oscillation is determined by integrating the time taken for a full cycle, given by T = \sqrt{2m} \int_{x_1}^{x_2} \frac{dx}{\sqrt{E - V(x)}}, where m is the particle mass and V(x) is the function. A prominent example of such bound motion occurs in parabolic potential wells, where the potential near its minimum approximates V(x) \approx \frac{1}{2} k x^2, leading to simple harmonic motion. In this case, the restoring force is linear, F = -k x, and the oscillation frequency is \omega = \sqrt{k/m}, independent of amplitude for small displacements. The period simplifies to T = 2\pi / \omega, providing an exact solution for the ideal harmonic oscillator. This approximation is widely applicable to systems like pendulums or molecular vibrations near equilibrium. Phase space analysis offers insight into these dynamics by plotting position x against momentum p = m \dot{x}. For integrable systems like the harmonic oscillator, trajectories form closed elliptical curves, reflecting periodic and predictable motion. In anharmonic wells, where V(x) deviates from quadratic (e.g., including cubic or quartic terms), phase space portraits reveal more complex structures; low-energy orbits remain quasi-periodic, but higher energies lead to chaotic trajectories filling irregular regions, as exemplified by the Hénon-Heiles potential V(x,y) = \frac{1}{2}(x^2 + y^2) + x^2 y - \frac{1}{3} y^3. While ideal conservative systems assume no energy loss, real-world trapped particles experience damping through dissipative forces like friction, causing amplitude decay and eventual equilibration at the well's minimum. However, analyses typically emphasize undamped cases to isolate intrinsic dynamics. Stability of these orbits is assessed via small perturbations around equilibrium; for linear wells, perturbations oscillate without growth, but nonlinear wells may exhibit instability quantified by positive Lyapunov exponents, indicating exponential divergence of nearby trajectories and the onset of chaos.

Stability and Escape Conditions

In classical mechanics, a particle remains trapped within a potential well if its total energy E is less than the height of the surrounding potential barrier, known as the escape energy E_\text{escape}. This condition ensures that the particle's kinetic energy becomes insufficient to surmount the barrier at its maximum, preventing escape to regions of higher potential. Particles with E \geq E_\text{escape} possess enough energy to classically overcome the barrier and propagate freely. In multi-dimensional or multi-well systems, such as double-well potentials, stability and escape are governed by saddle points that connect adjacent wells. The activation energy for transitioning between wells equals the potential energy at the saddle point relative to the minimum in the initial well; particles must reach this energy to cross the saddle and enter the neighboring well. For example, in the —a model for non-integrable escape dynamics—the threshold energy E_\text{th} = 1/6 marks the onset of possible escape from the central well over saddle points. Perturbations, including thermal fluctuations or external fields, can destabilize trapped particles by effectively lowering the barrier or providing transient energy boosts. In the presence of thermal noise, escape from a metastable well follows an activated process with a rate given by the Arrhenius-like expression \Gamma \propto \exp(-\Delta E / k_B T), where \Delta E is the barrier height, k_B is Boltzmann's constant, and T is temperature; this arises from for overdamped Brownian motion in a potential well. External fields, such as oscillating forces, similarly induce escape by modulating the potential landscape, with rates depending on the field's amplitude and frequency relative to the well's natural oscillation. For slowly varying potential wells, where parameters change adiabatically over timescales much longer than the particle's orbital period, the adiabatic invariant—the action variable J = \frac{1}{2\pi} \oint p \, dq—remains conserved. This conservation preserves the stability of bounded orbits, ensuring that the particle's motion adapts smoothly to the evolving well without escaping, even as the energy E adjusts proportionally. Classical stability criteria break down at high energies, where relativistic effects or chaotic dynamics dominate, and at quantum scales, where tunneling permits escape for E < E_\text{escape} despite the absence of sufficient classical energy. These limitations highlight the regime of validity for classical descriptions, typically applicable to macroscopic systems or high-energy quantum approximations.

Quantum Perspective

Wavefunctions and Energy Quantization

In quantum mechanics, bound states in a potential well arise as solutions to the time-independent Schrödinger equation for particle energies E below the height of the confining potential barriers. The equation is given by -\frac{\hbar^2}{2m} \frac{d^2 \psi(x)}{dx^2} + V(x) \psi(x) = E \psi(x), where \psi(x) is the wavefunction, V(x) is the potential energy profile, m is the particle mass, and \hbar is the reduced Planck's constant. For bound states, the solutions \psi_n(x) must be normalizable, meaning \int_{-\infty}^{\infty} |\psi_n(x)|^2 dx = 1, which confines the particle within the well and leads to discrete, quantized energy levels E_n with E_n < V_{\text{barrier}}. These eigenfunctions and eigenvalues form the stationary states of the system, describing particles trapped indefinitely in the absence of perturbations./02%3A_Resonances/2.01%3A_Bound_States_and_Free_States) The physical interpretation of these wavefunctions follows from Born's rule, which states that the probability density of finding the particle at position x is |\psi_n(x)|^2. This probabilistic nature reflects the intrinsic uncertainty in quantum mechanics, where the particle does not follow a definite trajectory but is delocalized according to the wavefunction's amplitude. Regions of high |\psi_n(x)|^2 indicate likely positions, while nodes—points where \psi_n(x) = 0—arise from quantum interference between de Broglie waves, resulting in zero probability density at those locations./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/07%3A_Quantum_Mechanics/7.02%3A_Wavefunctions) The lowest-energy solution, known as the ground state (n=1), features a wavefunction with no nodes (except possibly at boundaries) and is characterized by a single broad maximum centered in the well, maximizing overlap with the lowest potential regions. Higher excited states (n > 1) exhibit increasingly oscillatory wavefunctions with n-1 nodes, reflecting more rapid variations that accommodate higher energies while remaining confined. A key quantum feature is the zero-point energy, where the ground-state energy E_0 > 0 (relative to the classical minimum of the well), arising from the Heisenberg uncertainty principle, which prohibits the particle from being completely at rest even in the lowest state./03%3A_The_Schrodinger_Equation_and_a_Particle_in_a_Box/3.05%3A_The_Energy_of_a_Particle_in_a_Box_is_Quantized) In the semiclassical limit of deep and wide potential wells, the Bohr asserts that quantum energy levels E_n approximate classical allowed energies, with wavefunctions concentrating near classical turning points where E = V(x). This alignment is captured qualitatively by the Wentzel-Kramers-Brillouin (, which connects quantum quantization to classical action integrals without requiring exact solutions. For transitions between these quantized levels in harmonic-like potential wells, selection rules dictate allowed changes in the , typically \Delta n = \pm 1, as derived from the dipole approximation in time-dependent ; this restricts optical or vibrational excitations to adjacent states, ensuring conservation of and in the interaction ./06%3A_Vibrational_States/6.06%3A_Harmonic_Oscillator_Selection_Rules)

Tunneling Effects

In , a key feature distinguishing the quantum perspective from the classical one is the phenomenon of tunneling, where a particle with total energy E less than the height V of a potential barrier can penetrate into and traverse the classically forbidden region. This occurs because the particle's wavefunction, governed by the time-independent , does not abruptly terminate at the barrier's edge but instead decays evanescently in the forbidden region as \psi(x) \propto e^{-\kappa x}, where \kappa = \sqrt{2m(V - E)} / \hbar, m is the particle , and \hbar is the reduced Planck's constant. Although the probability density |\psi|^2 decreases exponentially, it remains finite, allowing a non-zero probability for the particle to emerge on the other side of the barrier. For a rectangular potential barrier of width L and height V, the transmission probability T—the ratio of transmitted to incident —can be approximated when \kappa L \gg 1 (thick barrier limit) as T \approx \exp\left(-2 \kappa L\right), where the exponential factor captures the dominant suppression due to the barrier's opacity. This approximation arises from matching wavefunctions across the barrier interfaces in the exact solution of the , neglecting oscillatory contributions inside the barrier for low E. More precise forms include a prefactor, but the exponential term sets the , highlighting how T decreases rapidly with increasing barrier width or height above E. In the context of potential wells, tunneling enables escape from bound states, with profound implications in various physical systems. A seminal application is in atomic nuclei, where an ( nucleus) tunnels through the surrounding the nuclear potential well. George Gamow's 1928 theory modeled the nucleus as a potential well confining the , predicting decay rates via the tunneling probability that matched experimental Geiger-Nuttall plots of decay constant versus alpha energy. This breakthrough explained why alpha particles with E far below the barrier height (~25 MeV) could escape, unifying radioactivity under . Similarly, in field emission from metals, electrons near the in the atomic potential wells tunnel through the surface image-potential barrier under a strong external , as described by the Fowler-Nordheim , enabling applications like electron microscopy. Resonant tunneling further illustrates barrier penetration in well structures, occurring in double-barrier configurations where two potential barriers sandwich a narrow . Here, transmission T is dramatically enhanced—approaching unity—when the incident particle's energy aligns with a quasi-bound state in the intermediate well, allowing coherent wavefunction overlap across both barriers. This arises from effects in the time-dependent solution, leading to sharp peaks in T(E). Experimentally observed in heterostructures like GaAs/AlGaAs, this effect was theoretically proposed and demonstrated by Esaki and Tsu, forming the basis for resonant tunneling diodes with negative differential resistance. While pure ground-state tunneling is temperature-independent, overall escape rates from potential wells often exhibit temperature dependence due to thermal population of excited states within the well. Higher-energy states have wavefunctions extending farther into the barrier, effectively reducing the tunneling and increasing T via smaller \kappa. At finite , the Boltzmann factor populates these states, yielding a net rate \Gamma(T) \propto \sum_n g_n e^{-E_n / kT} T_n, where g_n is degeneracy and T_n the from level n. This thermally assisted tunneling dominates in systems like molecular magnets or double wells at intermediate temperatures, bridging quantum and classical regimes.

Types and Models

Infinite Potential Well

The infinite potential well, also known as the infinite square well, is an idealized model in quantum mechanics where a particle of mass m is confined to a one-dimensional region between x = 0 and x = L, with the potential energy defined as V(x) = 0 for $0 < x < L and V(x) = \infty elsewhere. This setup enforces hard-wall boundary conditions, requiring the wavefunction to vanish at the boundaries: \psi(0) = \psi(L) = 0. The infinite barriers prevent the particle from existing outside the well, making the probability density zero beyond $0 \leq x \leq L. The time-independent Schrödinger equation within the well simplifies to -\frac{\hbar^2}{2m} \frac{d^2 \psi}{dx^2} = E \psi, yielding exact stationary state solutions for the wavefunctions and energies. The normalized energy eigenfunctions are \psi_n(x) = \sqrt{\frac{2}{L}} \sin\left( \frac{n \pi x}{L} \right), for n = 1, 2, 3, \dots, and the corresponding quantized energies are E_n = \frac{n^2 \pi^2 \hbar^2}{2 m L^2}. These solutions satisfy the boundary conditions and form orthonormal states, with the ground state (n=1) exhibiting a non-zero zero-point energy E_1 = \frac{\pi^2 \hbar^2}{2 m L^2}. Key properties of these solutions include the quadratic dependence of energy on the quantum number n, leading to energy spacings that increase with n: the difference \Delta E_n = E_{n+1} - E_n = \frac{(2n+1) \pi^2 \hbar^2}{2 m L^2}. The wavefunctions exhibit definite parity with respect to the well's center at x = L/2, alternating between even and odd symmetry: odd n yields even parity, while even n yields odd parity. Additionally, the set \{\psi_n(x)\} forms a complete orthonormal basis for square-integrable functions on [0, L], enabling the expansion of any valid wavefunction as \psi(x) = \sum_{n=1}^\infty c_n \psi_n(x) with \sum_{n=1}^\infty |c_n|^2 = 1. Despite its exact solvability, the model has limitations due to the unrealistic assumption of impenetrable infinite barriers, which neglects effects like tunneling in real systems. It remains a valuable pedagogical tool and approximation for quantum confinement in nanostructures, such as or wires, where barrier heights are high but finite.

Finite Potential Well

The finite potential well provides a more realistic model than the infinite well by allowing the confining potential to have finite depth, enabling wavefunction penetration into classically forbidden regions. The potential is defined as V(x) = -V_0 for |x| < a and V(x) = 0 elsewhere, where V_0 > 0 is the well depth and $2a is the width. Bound states occur for energies -V_0 < E < 0, while scattering states exist for E > 0. The time-independent yields oscillatory solutions inside the well and exponentially decaying solutions outside. Define k = \sqrt{2m(E + V_0)} / \hbar and \kappa = \sqrt{-2mE} / \hbar. For even-parity bound states, the wavefunction is \psi(x) = A \cos(kx) for |x| < a and \psi(x) = C e^{-\kappa |x|} for |x| > a; of the wavefunction and its at x = a leads to the \tan(ka) = \kappa / k. For odd-parity states, \psi(x) = B \sin(kx) inside and an odd extension outside, yielding -\cot(ka) = \kappa / k. These equations are solved graphically or numerically, with intersections determining the discrete allowed energies E_n. The number of bound states depends on the parameter proportional to V_0 a^2, specifically z_0 = a \sqrt{2m V_0} / \hbar; shallow or narrow wells support fewer states than the infinite well, and the highest energy approaches zero as the well becomes shallower. For example, increasing the well width from $2a_0 (one bound state) to $6a_0 (three bound states) illustrates this dependence. For scattering states with E > 0, the wavefunctions are plane waves outside the well with oscillatory behavior inside, forming a continuum of energies without quantization. An incident particle from the left has amplitude A for x < -a, transmitted amplitude F for x > a, and inside -a < x < a, \psi(x) = C \sin(\mu x) + D \cos(\mu x) where \mu = \sqrt{2m(E + V_0)} / \hbar. The transmission coefficient is T = |F/A|^2 = \left[ 1 + \frac{V_0^2 \sin^2(2 \mu a)}{4 E (E + V_0)} \right]^{-1}, which exhibits resonances where T = 1 at certain energies, akin to the Ramsauer-Townsend effect in electron scattering. The reflection coefficient is R = 1 - T. The one-dimensional finite well extends to three dimensions via the spherical symmetric potential V(r) = -V_0 for r < a and 0 otherwise, serving as a simple model for bound states in central potentials. For l = 0 (s-states), bound states require sufficient well depth and width, solved using the radial with spherical inside and modified exponentials outside, analogous to the 1D case but incorporating barriers.

Applications

Atomic and Molecular Physics

In atomic physics, the potential well governing electron behavior in the hydrogen atom is described by the Coulomb potential V(r) = -\frac{Z e^2}{r}, where Z is the atomic number, e is the elementary charge, and r is the radial distance from the nucleus. This attractive potential confines the electron, leading to quantized energy levels and wavefunctions solved via the Schrödinger equation, with the ground state energy at -13.6 eV for hydrogen (Z = 1). For multi-electron atoms, the shell structure arises from effective quantum wells formed by the nuclear attraction screened by inner electrons, resulting in an approximate central potential that organizes electrons into discrete shells labeled by principal quantum number n. These effective potentials explain the periodic table's filling order and chemical properties, as outer electrons experience a reduced nuclear charge due to electron-electron repulsion. In , potential wells form along the internuclear axis between atoms, dictating bond formation and stability. For van der Waals bonds, the models the weak attraction and repulsion between neutral atoms, capturing the equilibrium distance and for non-covalent interactions. Covalent bonds, in contrast, are described by the , which better approximates the of diatomic molecules, allowing for vibrational energy levels that decrease in spacing with increasing due to the well's finite depth. These vibrational levels within the molecular well enable the study of bond dynamics, where the prevents complete collapse to the minimum. Spectroscopy in atoms and molecules relies on transitions between quantized levels in these potential wells, producing absorption and emission lines corresponding to energy differences \Delta E. In diatomic molecules, the Franck-Condon principle governs the intensities of these vibronic transitions, as electronic jumps occur much faster than nuclear motion, favoring vertical transitions where the internuclear distance remains unchanged. This principle, combined with selection rules, explains the progression of spectral bands observed in molecular spectra. Ionization from atomic wells requires supplying exceeding the , such as 13.6 eV for hydrogen's electron, beyond which the particle escapes the confining potential. Quantum chemistry methods, such as the Hartree-Fock approximation, treat multi-electron systems by assuming each electron moves in an effective potential generated by the nucleus and the averaged field of other electrons, yielding self-consistent orbitals and energies. This mean-field approach provides a foundational description of atomic and molecular electronic structure, with improvements over exact solutions for heavier atoms by incorporating exchange effects to avoid unphysical electron pairing.

Solid-State Physics

In , potential wells play a fundamental role in describing behavior within crystalline materials. The periodic arrangement of cores in a generates a repeating potential landscape, where electrons are subject to a periodic potential V(r) that varies with the periodicity. This structure confines electrons into allowed energy states, as described by , which states that the wavefunctions take the form ψ(r) = e^{i \mathbf{k} \cdot \mathbf{r}} u(\mathbf{r}), with u(r) periodic and matching the . The resulting Bloch waves lead to the formation of energy bands, where allowed energies form continuous bands separated by forbidden band gaps, enabling the distinction between insulators, semiconductors, and metals based on band filling. This periodic confinement is essential for understanding electrical conductivity in solids, as electrons near band edges can be excited across gaps with minimal energy input in semiconductors. Defects and impurities disrupt the ideal periodicity, creating localized potential wells that introduce discrete energy levels within the band gaps. Vacancies or atoms form traps, while substitutional impurities like in act as shallow donors, producing potential wells approximately 45 meV below the conduction band edge that loosely bind extra electrons. Similarly, acceptors such as create shallow levels about 45 meV above the valence band, binding holes via analogous Coulombic wells. These levels, first systematically analyzed in the context of physics, enable controlled doping to tune carrier concentrations and facilitate p-n formation. The binding energies scale inversely with the constant and effective mass, making shallow levels prominent in wide-bandgap materials like GaAs. At the nanoscale, quantum confinement enhances these effects by shrinking the effective size of potential wells in structures like quantum dots and quantum wires. In quantum dots, such as CdSe nanocrystals, the finite size (typically 2–10 ) imposes three-dimensional confinement, approximating a particle-in-a-box model where levels discretize into atomic-like states, with the lowest transition increasing as the inverse square of the radius. This blueshift in and spectra, predicted by effective mass theory, allows size-tunable , as the confinement ħ²π²/(2μL²) dominates over bulk gaps for small L. Quantum wires extend this to one dimension, further quantizing the into subbands, which sharpens electrical and optical responses in nanowires. Excitons in semiconductors represent bound electron-hole pairs confined by a screened potential well, with the attractive interaction V(r) = -e²/(εr) forming hydrogen-like states. In materials like GaAs, Wannier excitons exhibit large Bohr radii (around 10–20 nm) due to high screening (ε ≈ 12), allowing delocalization over many sites while maintaining energies of 4–10 meV at temperature.31002-6) These composite quasiparticles mediate efficient radiative recombination, influencing efficiency and contributing to the indirect-to-direct bandgap in some contexts. In electronic devices, engineered potential wells enable precise carrier confinement to boost performance. In MOSFETs, the gate-induced electric field forms a triangular potential well at the oxide-semiconductor , confining inversion-layer electrons to a ~10 nm thick , which reduces scattering and enhances mobility. For , multiple quantum wells in LEDs, such as InGaN/GaN heterostructures, trap electrons and holes in thin (2–5 nm) layers, increasing recombination probability and radiative efficiency by over an compared to bulk devices. These confinement strategies underpin modern high-speed transistors and efficient light emitters.

Nuclear and Particle Physics

In , the potential well describing the interaction between nucleons within atomic nuclei is often modeled using the Woods-Saxon potential, which provides a smooth approximation to the nuclear surface. This potential takes the form V(r) \approx -\frac{V_0}{1 + \exp\left(\frac{r - R}{a}\right)}, where V_0 is the depth (typically around 50 MeV), R is the nuclear radius proportional to A^{1/3} (with A the ), and a is the surface diffuseness parameter (about 0.5–0.7 fm). This form captures the binding of nucleons in a finite well, leading to quantized energy levels in the , where protons and neutrons occupy discrete orbitals analogous to shells in atoms, explaining like 2, 8, 20, and 28. Alpha clustering represents a substructure within heavier nuclei, where groups of four nucleons form alpha particles ( nuclei) that occupy localized potential wells inside the overall potential. This phenomenon is prominent in light and medium-mass nuclei, such as ^{12}\mathrm{C} and ^{16}\mathrm{O}, and is modeled by semi-microscopic approaches that incorporate alpha-alpha interactions as secondary wells, enhancing stability against or . Experimental evidence from resonant and reactions supports these clustered configurations, which lower the ground-state energy compared to independent models. In , the concept of potential wells extends to quark confinement within , governed by (QCD). The strong force between rises linearly with separation, modeled as V(r) \approx kr (with string tension k \approx 1 GeV/fm), creating an effectively infinite well that prevents free from being observed, as the energy required to separate them exceeds the threshold for creating new quark-antiquark pairs. simulations confirm this confining potential, reproducing masses and spectra without at long distances. Beta decay involves the enabling a within the potential well, converting a to a proton (or vice versa) while emitting an and antineutrino. This process occurs between bound states of the and nuclei, with half-lives determined by the overlap of wavefunctions and the Gamow-Teller , as seen in decays like ^{14}\mathrm{C} \to ^{14}\mathrm{N}. For deep potential wells in heavy nuclei, relativistic effects become significant, necessitating modifications to the to account for the strong scalar and vector fields. The Dirac phenomenology incorporates a large (\approx -400 MeV) and (\approx +350 MeV), yielding a small spin-orbit splitting that matches observed single-particle levels in nuclei like lead-208, unlike non-relativistic models. These relativistic mean-field theories, such as the Walecka model, provide accurate binding energies and radii for superheavy elements.

References

  1. [1]
    [PDF] Bound States in One Dimension
    By a potential well, we mean a graph of potential energy as a function of coordinate x. In this well picture, we indicate a constant energy level (total.
  2. [2]
    [PDF] Quantum Physics I, Lecture Note 11 - MIT OpenCourseWare
    Mar 17, 2016 · Here we introduce another instructive toy model, the infinite square well potential. This forces a particle to live on an interval of the real ...
  3. [3]
    [PDF] Lecture 12 The Finite Potential Well: A Quantum Well
    The Finite Potential Well: A Quantum Well. In this lecture you will learn: • Particle in a finite potential well. • Bound and unbound states in quantum physics ...Missing: definition | Show results with:definition
  4. [4]
    [PDF] Exact solutions of the quantum double square well potential
    For a symmetrical quantum double square well potential, we find analytical expressions satisfied by the quantized energies. Graphical or numerical solutions of ...
  5. [5]
    [PDF] 41.6 – Finite Potential Wells
    The diameter of the potential well is equal to the diameter of the nucleus (this varies with atomic mass), and nuclear physics experiments have found that the ...
  6. [6]
    [PDF] Optical Physics of Quantum Wells - Stanford Electrical Engineering
    Both electrons and holes see lower energy in the "well" layer, hence the name (by analogy with a "potential well"). This layer, in which both electrons and ...
  7. [7]
  8. [8]
    Gravitational Potential Energy - HyperPhysics
    Gravitational potential energy is energy an object possesses because of its position in a gravitational field.Missing: well | Show results with:well
  9. [9]
    Electron states in atoms - Benjamin Klein
    The electron waves in a Hydrogen atom (the simplest atom) are trapped in a 3d spherically-symmetric potential well. Describing the energy states of electrons ...<|control11|><|separator|>
  10. [10]
    [PDF] A Brief History of The Development of Classical Electrodynamics
    1828: George Green generalizes and extends the work of Lagrange, Laplace, and Poisson and attaches the name potential to their scalar function.
  11. [11]
    7. Electrostatics II: Conductors, Green's Theorem, Green's Functions
    The Story So Far… After Coulomb determined the inverse square law of electrical force in 1785, his French theoretical colleagues Lagrange, Laplace, Poisson and ...
  12. [12]
    E1-12: MARBLE IN GLASS BOWL - lecdem.physics.umd.edu
    Jun 13, 2014 · Purpose: Demonstrate orbits in a gravitational potential well. Description: Start the marble rolling tangentially in the bowl to obtain an orbit ...
  13. [13]
    15.2 Energy in Simple Harmonic Motion - UCF Pressbooks
    Consider the marble in the bowl example. If the bowl is right-side up, the ... Note that unlike the simple harmonic oscillator, the potential well of the Lennard- ...
  14. [14]
    [PDF] The Physics of Quantum Mechanics
    ... marble, then you're visualizing the atom incorrectly. Once again, our ... bowl it comes to rest at the deepest point and spends most of its time there ...<|control11|><|separator|>
  15. [15]
    Secondary school students' misunderstandings of potential wells ...
    Jun 1, 2020 · The misconception “potential wells are objects,” can be related to the phrase “well” and therefore may be caused by a linguistic impediment. ...Missing: common | Show results with:common
  16. [16]
    [PDF] Central Force Motion† 1. Introduction
    This is an example of central force motion because the constraint r = r0 can be thought of as due to a sharply confining potential V (r) centered on r = r0. r0.
  17. [17]
    [PDF] Notes on (calculus based) Physics - K. V. Shajesh
    Jan 18, 2021 · extremum points in the potential energy profile, where the force (given by the slope) is zero. An extremum point x0 is a stable point, an ...
  18. [18]
    [PDF] Chapter 7 The Schroedinger Equation in One Dimension In classical ...
    For a rounded well the potential energy U(x) changes continuously. We will assume that the rounded well has U(x) = U0 for x < 0 and x>a as shown in Figure 8.
  19. [19]
    [PDF] Chapter 8 The Simple Harmonic Oscillator
    The functional form of a simple harmonic oscillator from classical mechanics is V (x) = 1 2 kx2 . Its graph is a parabola as seen in the figure on the left.
  20. [20]
    [PDF] Regularized semiclassical radial propogator for the Coulomb potential
    the standard Van Vleck form for the unmodified Coulomb potential V(x) = —Z/x: The classical mechanicsfor the one-dimensional Coulomb problem with elastic ...
  21. [21]
    [PDF] The Morse Potential∗
    Feb 13, 2003 · The Morse potential is the simplest representative of the potential between two nuclei in which dissociation is possible. r0 x. U(x) equi−spaced.Missing: functional physics
  22. [22]
    [PDF] Finite Square Well Bound States R.M. Suter Preliminaries 1. First ...
    ǫ0 depends on the shape of the well: the depth, V0, and the width, a, through the product,. V0a2. This parameter sets the “strength” of the well and the number ...
  23. [23]
    [PDF] VII. Central Potentials
    Spherical Polar Coordinates. Since we are dealing with a potential that is a ... coordinates where r is one of the basic variables, rather than some.Missing: physics | Show results with:physics
  24. [24]
    [PDF] 1. Newtonian Mechanics
    ... potential V(x), is described in terms of an Equation of Motion relating the parti- cle's momentum p to the gradient of the potential. Fx = - {grad V(x)}x = V ...
  25. [25]
    6. Hamilton's Equations - Galileo and Einstein
    we have immediately the so-called canonical form of Hamilton's equations of motion: ∂H∂pi=˙qi,∂H∂qi=−˙pi. Evidently going from state space to phase space has ...<|separator|>
  26. [26]
    [PDF] The Schrödinger Equation in One Dimension
    Let us now apply the TISE to a simple system - a particle in an infinitely deep potential well. Particle in a One-Dimensional Rigid Box (Infinite Square Well).
  27. [27]
    [PDF] 1 The Schrödinger equation - MIT OpenCourseWare
    Sep 13, 2013 · Any of the three boxed equations above is referred to as the time-independent Schrödinger equation. Since the time-independent Schrödinger ...<|control11|><|separator|>
  28. [28]
    [PDF] 07. Lagrangian Mechanics III - DigitalCommons@URI
    Nov 5, 2015 · Find the value of a which makes the period of oscillation one second (1s) for g = 9.8m/s2 ... which the action integral is an extremum.
  29. [29]
    [PDF] Classical Mechanics - Rutgers Physics
    Nov 17, 2010 · One view of classical mechanics is as a steepest path approximation to the path integral which describes quantum mechanics. ... potential well ...
  30. [30]
    [PDF] Chapter 23 Simple Harmonic Motion - MIT OpenCourseWare
    Jul 23, 2022 · The mechanical energy in state 2 is equal to the initial potential energy in state 1, so the mechanical energy is constant. This should come as ...
  31. [31]
    [PDF] Jürgen Vollmer: Theoretical Mechanics - Universität Leipzig
    Figure 6.9 shows the effective potential and phase space portraits for different values of angular momentum, i. e. of the dimension- less control parameter ...
  32. [32]
    [PDF] Chaos in the Hénon-Heiles system
    This potential is now known as the Hélion-Heiles potential. It can be seen as two harmonic oscillators that has been coupled by the perturbation terms x2 y ...
  33. [33]
    Efficient numerical calculation of Lyapunov-exponents and stability ...
    Apr 5, 2024 · Investigating the stability of stationary motions is a highly relevant aspect when characterizing dynamical systems.
  34. [34]
    Particle in Finite Square Potential Well
    ... classical mechanics, a particle whose overall energy, $E$ , is negative is bound in the well (i.e., it cannot escape to infinity), whereas a particle whose ...Missing: threshold | Show results with:threshold
  35. [35]
    Threshold law for escaping from the Hénon-Heiles system
    Aug 22, 2007 · E th = 1 ∕ 6 is the threshold energy of this system. When E ⩾ E th , a particle in the potential well can escape. Figure 1 shows the ...
  36. [36]
    [PDF] THE KRAMERS PROBLEM: FIFTY YEARS OF DEVELOPMENT
    Kramers considered the escape rate of a Brownian particle out of a deep potential well. His final results are applicable in the regimes of strong and extremely ...
  37. [37]
    7.6 The Quantum Tunneling of Particles through Potential Barriers
    Sep 29, 2016 · Quantum tunneling is a phenomenon in which particles penetrate a potential energy barrier with a height greater than the total energy of the ...
  38. [38]
    [PDF] Chapter 4 Time–Independent Schrödinger Equation
    For n = 1 we get the ground state energy and wave function E1,ψ1 of the infinite potential well, the higher states with n > 1 are called excited states.
  39. [39]
    [PDF] 4. The Infinite Square Well - Physics
    But this model is still useful for understanding a variety of real-world potential wells (e.g., electrons in long organic molecules, or in fabricated.
  40. [40]
    [PDF] The WKB Method† 1. Introduction - University of California, Berkeley
    We will begin with the time-independent Schrödnger equation in three dimensions for a particle moving in a potential V (x), because the multidimensional case.<|control11|><|separator|>
  41. [41]
  42. [42]
    [PDF] Quantum Physics III Chapter 3: Semiclassical Approximation
    Feb 3, 2019 · The exact formula for the tunneling probability of a square barrier has been calculated before, and it is given by. 1. T. = 1+. V 2. 0. 4E(V0 − ...
  43. [43]
    The Quantum Theory of Nuclear Disintegration - Nature
    The Quantum Theory of Nuclear Disintegration. G. GAMOW. Nature volume 122, pages 805–806 (1928)Cite this article. 1329 Accesses. 103 Citations. 3 Altmetric.
  44. [44]
    Electron emission in intense electric fields - Journals
    Lepetit B (2017) Electronic field emission models beyond the Fowler-Nordheim one, Journal of Applied Physics, 10.1063/1.5009064, 122:21, Online publication ...
  45. [45]
    Resonant tunneling in semiconductor double barriers - AIP Publishing
    Jun 15, 1974 · Esaki, R. Tsu; Resonant tunneling in semiconductor double barriers. Appl. Phys. Lett. 15 June 1974; 24 (12): 593–595. https://doi.org ...
  46. [46]
    Quantum tunnelling and thermally driven transitions in a double-well ...
    May 29, 2024 · We explore dissipative quantum tunnelling, a phenomenon central to various physical and chemical processes, utilizing a model based on a double-well potential.<|control11|><|separator|>
  47. [47]
    Particle in Infinite Square Potential Well
    The most general wavefunction for a particle trapped in a one-dimensional square potential well, of infinite depth, is a superposition of all of the possible ...
  48. [48]
    [PDF] Infinite (and finite) square well potentials - High Energy Physics
    x. 0 a. V(x). E. Infinite square well approximation assumes that electrons never get out of the well so V(0)=V(a)=∞ and ψ(0)=ψ(a)=0. A more accurate potential.
  49. [49]
    None
    ### Summary of Scattering States in Finite Square Well
  50. [50]
    [PDF] Finite Square Well - UNCW
    If one were to increase the with of the potential well to a = 6a0, then there would be two bound states. The graphical solution of transcendental equation is ...
  51. [51]
    [PDF] 4 A Quantum Particle in Three Dimensions - DAMTP
    Consider the spherical finite potential well,. V (r) = (. V0 r<a. 0 otherwise ... In the context of the hydrogen atom it is sometimes called the principal quantum ...
  52. [52]
    [PDF] The Hydrogen Atom: a Review on the Birth of Modern Quantum ...
    The purpose of this work is to retrace the steps that were made by scientists of. XXcentury, like Bohr, Schrodinger, Heisenberg, Pauli, Dirac, ...
  53. [53]
    Quantum Mechanics of the Hydrogen Atom | SpringerLink
    In this chapter, we shall solve the Schrödinger equation of the hydrogen atom. For our calculations, we will not initially restrict ourselves to the Coulomb ...Missing: paper | Show results with:paper
  54. [54]
    [PDF] Elements of Atomic Structure in Multi-Electron Atoms
    When we obtain the Hartree-Fock estimate for the energy, the result will depend only on the quantum numbers (nℓ), and not on the magnetic quantum numbers (mℓ,ms) ...
  55. [55]
    [PDF] Multielectron Atoms
    Within the IPA, each of the Z electrons in an atom moves independently with a spherically symmetric net potential energy U(r) due to its attraction to the ...
  56. [56]
    On the determination of molecular fields. —II. From the equation of ...
    The object of the present paper is to ascertain whether a molecular model of the same type will also explain available experimental data concerning the equation ...
  57. [57]
    Diatomic Molecules According to the Wave Mechanics. II. Vibrational ...
    An exact solution is obtained for the Schroedinger equation representing the motions of the nuclei in a diatomic molecule.
  58. [58]
    (PDF) Morse potential derived from first principles - ResearchGate
    Aug 7, 2025 · ... Article. Jan 1929. Philip M. Morse. An exact solution is obtained for the Schroedinger equation representing the motions of the nuclei in a ...
  59. [59]
    [PDF] About the Quantum Mechanics of the Electrons in Crystal Lattices
    Nov 29, 2018 · Abstract: This article aims to review Felix Bloch theorem of electron motion in a crystal lattice through his seminal paper that has also ...
  60. [60]
    First-principles calculations for point defects in solids | Rev. Mod. Phys.
    Mar 28, 2014 · For a shallow donor or acceptor, the ionization energy determines the fraction of dopants that will be ionized and hence contribute free ...
  61. [61]
    [PDF] Electrons and Holes in Semiconductors - Pearl HiFi
    Shockley leads was aimed at under standing a kind of materials, the semiconductors, which had already received considerable application in the ...
  62. [62]
    [PDF] The size dependence of the lowest excited electronic state - FKIT
    Nov 8, 2011 · Electron-electron and electron-hole interactions in small semiconductor ... L. E. Brus: Interactions in small semiconductor crystallites. 4405.