Fact-checked by Grok 2 weeks ago

Rectangular potential barrier

In , the rectangular potential barrier, also known as the square potential barrier, is a one-dimensional model potential defined as V(x) = 0 for x < 0 and x > L, and V(x) = V_0 (a constant) for $0 \leq x \leq L, where L is the barrier width and V_0 > 0 is its height. This idealized setup represents a finite barrier that a particle of mass m and E encounters, dividing space into three regions: incident/reflected waves before the barrier, evanescent or propagating waves inside, and transmitted waves after. The time-independent is solved in each region, with continuity of the wave function \psi(x) and its derivative \psi'(x) enforced at the boundaries x = 0 and x = L to determine the coefficients. For E > V_0, the solutions are oscillatory plane waves in all regions, with wave numbers k_1 = \sqrt{2mE}/\hbar outside and k_2 = \sqrt{2m(E - V_0)}/\hbar inside the barrier, leading to partial reflection and transmission analogous to classical waves at an . The R and T (with R + T = 1) are given by T = \frac{4 k_1^2 k_2^2}{4 k_1^2 k_2^2 + (k_1^2 - k_2^2)^2 \sin^2(k_2 L)} and R = 1 - T, exhibiting resonances where T = 1 at certain energies when k_2 L = n\pi (integer n). In contrast, for $0 < E < V_0, the wave function decays exponentially inside the barrier with decay constant \kappa = \sqrt{2m(V_0 - E)}/\hbar, while remaining oscillatory outside, enabling quantum tunneling where the transmission probability T \approx 16 \frac{E}{V_0} (1 - \frac{E}{V_0}) e^{-2\kappa L} for thick barriers (\kappa L \gg 1), defying classical expectations of total reflection. The exact T is T = \left[1 + \frac{V_0^2 \sinh^2(\kappa L)}{4E(V_0 - E)}\right]^{-1}, highlighting the exponential sensitivity to barrier width and height. This model is foundational for understanding scattering processes and non-classical penetration in quantum systems, with applications in phenomena like alpha decay, scanning tunneling microscopy, and semiconductor device physics, where tunneling currents enable nanoscale imaging and electron transport. The scattering is often described using the unitary S-matrix, relating incoming and outgoing wave amplitudes, ensuring probability conservation.

Introduction

Physical setup

The rectangular potential barrier serves as a canonical model in one-dimensional for studying particle scattering and transmission through a localized potential obstacle. It features a finite region of elevated constant potential energy flanked by areas of zero potential, enabling analysis of both classical and quantum behaviors in wave propagation. The potential energy is piecewise defined as V(x) = \begin{cases} 0 & x < 0 \\ V_0 & 0 \leq x \leq a \\ 0 & x > a \end{cases} where V_0 > 0 denotes the barrier height and a > 0 its width. A quantum particle, such as an , approaches the barrier from the left in the region x < 0 with total energy E, which may satisfy E < V_0, E = V_0, or E > V_0. This setup divides the space into three distinct regions: Region I (x < 0) for the incident and possible reflected waves, Region II ($0 \leq x \leq a) inside the barrier, and Region III (x > a) for any transmitted wave. Classically, a particle with E < V_0 is forbidden from entering the barrier and reflects completely at the edge; at E = V_0, it reaches the barrier boundary but lacks sufficient energy to transmit; and for E > V_0, it passes through unimpeded with no . In contrast, reveals non-classical effects due to the wave-like nature of particles: tunneling allows a finite probability even when E < V_0, while partial reflection occurs for E \geq V_0 due to wave interference at the discontinuities. These differences underscore the barrier's role in illustrating quantum penetration and scattering phenomena. Visually, the potential V(x) is represented as a plot against x, depicting a sharp rectangular step rising uniformly to height V_0 from x = 0 to x = a, with flat zero levels extending to negative and positive infinity. The particle's energy E appears as a dashed horizontal line across the plot, positioned below, at, or above V_0 to indicate the relevant regime, with vertical boundaries marking the three regions where the wave function exhibits oscillatory or evanescent behavior. This physical configuration provides the foundation for applying the time-independent Schrödinger equation to determine the wave functions and transmission properties in each energy case.

Importance in quantum mechanics

The rectangular potential barrier emerged as a key model in the formative years of quantum mechanics during the late 1920s, when researchers sought to explain phenomena involving particle penetration through energy barriers. Pioneering work by Friedrich Hund in 1927 applied barrier penetration concepts to molecular spectra, while Lothar W. Nordheim that same year employed rectangular barriers to model electron emission from metals under electric fields, providing one of the earliest explicit uses of this idealized geometry. George Gamow extended these ideas in 1928 to interpret alpha decay in radioactive nuclei, using tunneling through a barrier—though his potential was more realistically Coulombic, the rectangular form served as a foundational simplification for such calculations. This model delivers a profound insight into quantum behavior by illustrating how particles can traverse regions of space where their total energy is less than the potential energy, a process utterly forbidden in classical physics. Unlike classical particles, which would reflect completely from such an obstacle, quantum particles exhibit a finite transmission probability, underscoring the probabilistic essence of quantum mechanics and defying everyday intuitions about impenetrable walls. Central to this is the manifestation of wave-particle duality, where the particle's associated wave function does not abruptly terminate at the barrier's edge but exponentially decays into the forbidden region, ensuring a non-vanishing amplitude on the far side and thus enabling probabilistic transmission. This penetration highlights the wavelike character of matter, first hypothesized by , and reveals how quantum waves interfere constructively or destructively to determine outcomes like reflection or transmission. As a pedagogical cornerstone, the rectangular barrier represents the simplest exactly solvable case of potential scattering in one dimension beyond the free-particle scenario, allowing exact analytic solutions that illuminate without the complexities of more realistic potentials. Its tractability has made it indispensable in quantum mechanics curricula, fostering conceptual grasp of non-intuitive effects like evanescent waves and boundary matching.

Mathematical model

Definition of the potential

The rectangular potential barrier is a canonical model in one-dimensional quantum mechanics used to investigate particle scattering and quantum tunneling phenomena. It consists of a finite region of constant positive potential energy separating two regions of zero potential, approximating idealized barriers encountered in atomic and nuclear physics. The potential is mathematically defined as a piecewise function along the x-axis: V(x) = \begin{cases} 0 & x < 0 \quad (\text{region I}) \\ [V_0](/page/Height) & 0 \leq x \leq a \quad (\text{region II}) \\ 0 & x > a \quad (\text{region III}) \end{cases} where V_0 > 0 represents the barrier (in units), and a > 0 denotes the barrier width (in units). The incident particle E satisfies $0 < E, typically with analysis considering cases where E < V_0 (tunneling regime) or E > V_0 (above-barrier ). This setup assumes non-relativistic motion confined to one along the x-direction, with the potential symmetric in height across the barrier but asymmetric due to unidirectional incidence from the left (region I). In theoretical derivations, natural units are commonly adopted for simplicity, setting the reduced Planck's constant \hbar = 1 and particle mass m = 1, which scales energies in units of \hbar^2 / (m a^2) and lengths in units of a. This potential form directly substitutes into the time-independent Schrödinger equation to yield region-specific wave equations, as explored in subsequent analysis.

Time-independent Schrödinger equation

The time-independent describes the stationary states of a quantum particle in the rectangular potential barrier, assuming a time-independent \psi(x) for scattering problems. The full is then \Psi(x,t) = \psi(x) e^{-iEt/\hbar}, where the spatial part \psi(x) is normalized such that the is constant, reflecting the steady-state nature of the incident, reflected, and transmitted fluxes. In its general one-dimensional form, the equation is -\frac{\hbar^2}{2m} \frac{d^2 \psi(x)}{dx^2} + V(x) \psi(x) = E \psi(x), where \hbar is the reduced Planck's constant, m is the mass of the particle, E > 0 is the total energy, and V(x) is the potential energy. This can be rearranged as \frac{d^2 \psi(x)}{dx^2} = \frac{2m}{\hbar^2} \left[ V(x) - E \right] \psi(x). The setup considers a plane wave incident from the left, so the asymptotic behavior in the leftmost region is \psi(x \to -\infty) \sim e^{ikx} + r e^{-ikx}, where r is the complex reflection amplitude and k is the wave number defined below. The potential V(x) divides space into three regions: regions I and III where V(x) = 0 (outside the barrier), and region II where V(x) = V_0 > 0 (inside the barrier). In regions I and III, the equation reduces to the free-particle form \frac{d^2 \psi}{dx^2} + k^2 \psi = 0, with the wave number k = \sqrt{2mE}/\hbar. In region II, the form depends on the energy regime relative to the barrier height V_0. For sub-barrier energies (E < V_0), it becomes \frac{d^2 \psi}{dx^2} - \kappa^2 \psi = 0, where the decay constant is \kappa = \sqrt{2m(V_0 - E)}/\hbar. For above-barrier energies (E > V_0), the equation is instead \frac{d^2 \psi}{dx^2} + q^2 \psi = 0, with the wave number q = \sqrt{2m(E - V_0)}/\hbar. These region-specific equations capture the oscillatory behavior outside the barrier and the evanescent or oscillatory character inside, setting the stage for solving the boundary-value problem.

Wave functions in different regions

Incident and reflected waves (region I)

In region I, where x < 0 and the potential V(x) = 0, the time-independent Schrödinger equation yields oscillatory solutions for the wave function, representing free-particle propagation. The general form of the wave function is \psi_I(x) = A e^{i k x} + B e^{-i k x}, where A is the amplitude of the incident , B is the amplitude of the reflected , and the wave number k = \sqrt{2 m E}/\hbar, with m the particle mass, E the energy of the incident particle, and \hbar the reduced Planck's constant. The term A e^{i k x} describes the incident wave propagating to the right toward the barrier, while B e^{-i k x} represents the reflected wave propagating to the left away from the barrier. This superposition captures the scattering process where the particle approaches from -\infty and part of the wave is backscattered. Often, for normalization in scattering problems, the incident amplitude is set to A = 1, simplifying calculations while preserving the ratios of amplitudes. The probability current in region I ensures conservation of probability flux, with the incident current given by j_{\text{inc}} = (\hbar k / m) |A|^2, directed to the right, and the reflected current j_{\text{ref}} = -(\hbar k / m) |B|^2, directed to the left. Asymptotically, as x \to -\infty, the wave function consists solely of the incident wave incoming from the left with no additional incoming component from -\infty, consistent with the scattering setup.

Transmitted wave (region III)

In region III, defined for x > L where the potential V(x) = 0, the time-independent yields the same general solution form as in region I, consisting of plane waves with wave number k = \sqrt{2mE}/\hbar. The appropriate wave function, assuming particle incidence solely from the left and no incoming wave from the right, excludes the left-propagating component, resulting in \psi_{\text{III}}(x) = C e^{ikx}, where C is a complex amplitude constant. This form describes a purely transmitted right-moving wave, carrying the portion of the incident probability flux that penetrates the barrier. The magnitude |C| governs the transmission probability, defined relative to the incident wave amplitude as T = |C|^2 / |A|^2, where A is the coefficient of the incident wave in region I. The associated probability current density, representing the transmitted particle flux, is given by j_{\text{trans}} = \frac{\hbar k}{m} |C|^2. This expression follows from the general formula for in one dimension and quantifies the rate at which probability density flows to the right in region III. As x \to \infty, the wave function behaves as \psi_{\text{III}}(x) \sim e^{ikx} up to an overall in C, indicative of propagation. For incident energies well above the barrier height (E \gg V_0), this asymptotic form corresponds to near-complete transmission, where |C| \approx |A| and reflection becomes negligible.

Evanescent or oscillatory wave (region II)

In the region inside the rectangular potential barrier (region II, where $0 < x < L and V(x) = V_0), the time-independent Schrödinger equation yields general solutions for the wave function \psi_{II}(x) that depend on the relationship between the particle's energy E and the barrier height V_0. These solutions take the form of linear combinations of basis functions, with coefficients determined by boundary conditions at the interfaces. For sub-barrier energies where E < V_0, the solutions are evanescent waves, expressed as \psi_{II}(x) = D e^{\kappa x} + F e^{-\kappa x}, with the decay constant \kappa = \sqrt{2m(V_0 - E)} / \hbar. Here, m is the particle mass and \hbar is the reduced Planck's constant. The term e^{-\kappa x} represents exponential decay propagating to the right, while e^{\kappa x} represents growth, illustrating the classically forbidden nature of the region where the wave function penetrates the barrier via without classical oscillation. This evanescent behavior ensures the wave function remains finite and continuous, avoiding unphysical divergences. For above-barrier energies where E > V_0, the solutions become oscillatory waves, given by \psi_{II}(x) = D e^{i q x} + F e^{-i q x}, where the wave number q = \sqrt{2m(E - V_0)} / \hbar. The e^{i q x} term corresponds to a rightward-propagating wave, and e^{-i q x} to a leftward-propagating one, reflecting partial inside the barrier similar to free-particle motion but with a reduced effective E - V_0. This oscillatory form allows for classical transmission over the barrier, modulated by quantum reflection effects. In both cases, the general form \psi_{II}(x) is a superposition ensuring of the wave function and its at x = 0 and x = L, with coefficients D and F (or equivalent labels) solved via matching to the exterior regions. At the E = V_0, the solutions simplify to linear s, \psi_{II}(x) = D + F x, marking the transition between evanescent and oscillatory regimes.

Derivation of transmission and reflection

Boundary condition matching

To solve for the unknown coefficients in the wave functions describing the rectangular potential barrier, the continuity of the wave function \psi(x) and its first derivative \psi'(x) must be enforced at the boundaries where the potential changes abruptly. These boundary conditions arise from the requirement that the time-independent Schrödinger equation yields a well-behaved, finite solution for a finite potential step, ensuring no infinite discontinuities in the probability density or current. For the standard setup with the barrier extending from x = 0 to x = a, the interfaces occur at x = 0 and x = a. Assuming the wave function forms from the regions as previously outlined—incident and reflected plane waves in region I (x < 0), evanescent waves in region II ($0 < x < a) for sub-barrier energies, and transmitted plane waves in region III (x > a)—the matching is applied with the incident amplitude normalized to unity (A = 1). At x = 0, continuity of \psi(x) gives: $1 + B = D + F, where B is the reflection amplitude, and D and F are the coefficients of the growing (e^{\kappa x}) and decaying (e^{-\kappa x}) terms in region II, respectively, with \kappa = \sqrt{2m(V_0 - E)} / \hbar for E < V_0. Continuity of \psi'(x) yields: ik(1 - B) = \kappa(D - F), where k = \sqrt{2mE} / \hbar. Similar conditions apply at x = 0 for E > V_0, replacing \kappa with the oscillatory wave number q = \sqrt{2m(E - V_0)} / \hbar and adjusting the derivative signs accordingly. At x = a, continuity of \psi(x) requires: D e^{\kappa a} + F e^{-\kappa a} = C e^{i k a}, where C is the transmission amplitude in region III. The derivative continuity condition is: \kappa \left( D e^{\kappa a} - F e^{-\kappa a} \right) = i k C e^{i k a}. Again, for E > V_0, the forms adapt to oscillatory waves in region II, with \kappa replaced by i q. These four equations form a linear system that can be expressed in matrix form for the unknowns B, C, D, and F. The resulting 4×4 coefficient matrix ensures the probabilistic current is conserved across the barrier, satisfying |A|^2 = |B|^2 + |C|^2 due to the identical wave numbers k in regions I and III, which maintains unitarity of the scattering matrix.

Expressions for coefficients

The reflection and transmission coefficients for the rectangular potential barrier are derived from the amplitudes B and C of the reflected and transmitted , respectively, obtained via boundary condition matching at the potential interfaces. Assuming unit incident amplitude (A = 1), the is R = |B|^2 and the is T = |C|^2, with the relation R + T = 1 holding due to conservation of . For sub-barrier energies (E < V_0), where a is the full width of the barrier, the corresponding transmission coefficient is T = \left[ 1 + \frac{V_0^2 \sinh^2(\kappa a)}{4 E (V_0 - E)} \right]^{-1}, and R = 1 - T. For above-barrier energies (E > V_0), the expressions are obtained by analytic continuation: replace \kappa \to i q where q = \sqrt{2 m (E - V_0)}/\hbar, and \sinh(\kappa a) \to i \sin(q a). This yields an oscillatory transmission coefficient T = \left[ 1 + \frac{V_0^2 \sin^2(q a)}{4 E (E - V_0)} \right]^{-1}, with R = 1 - T, reflecting resonant transmission when \sin(q a) \approx 0. A general expression for the , valid across energy regimes via , is T = \left[ 1 + \frac{(k^2 + \kappa^2)^2 \sinh^2(\kappa a)}{4 k^2 \kappa^2} \right]^{-1}, where k^2 + \kappa^2 = 2 m V_0 / \hbar^2, reducing to the specific forms above for E < V_0 and E > V_0.

Analysis by energy regimes

Sub-barrier energies (E < V₀)

When the particle energy E is less than the barrier height V_0, classical mechanics predicts complete reflection at the barrier interface, as the particle lacks sufficient kinetic energy to surmount it. In quantum mechanics, however, there is a non-zero probability for the particle to tunnel through the barrier due to the evanescent wave in region II, where the wave function decays exponentially rather than oscillating. This tunneling probability, derived from matching boundary conditions on the exact transmission coefficient, highlights a purely quantum effect with no classical counterpart. For thick barriers where \kappa L \gg 1, with \kappa = \sqrt{2m(V_0 - E)} / \hbar and L the full barrier width, the transmission probability T approximates to T \approx \frac{16 E (V_0 - E)}{V_0^2} e^{-2 \kappa L}, demonstrating an exponential decay with increasing barrier width L, modulated by a prefactor that peaks near E = V_0 / 2. This approximation arises from the hyperbolic sine term in the exact expression dominating for large arguments, emphasizing the sensitivity of tunneling to barrier dimensions and energy deficit. The corresponding reflection probability is R \approx 1 - T, indicating nearly total reflection but with a small, finite transmission that enables observable quantum effects in applications like . In the one-dimensional model, transmission is independent of any "angle of incidence" since motion is collinear, but T decreases rapidly as the barrier width increases, underscoring the evanescent nature of the wave inside the barrier. Unlike infinitely thin delta-function barriers, which exhibit energy-dependent transmission without perfect transmission peaks, the finite-width rectangular barrier in this regime shows no resonances or perfect transmission points; the probability monotonically decreases without oscillatory features.

Above-barrier energies (E > V₀)

When the energy E of an incident particle exceeds the barrier height V_0, predicts complete transmission without reflection, as the particle surmounts the potential. However, quantum mechanically, the wave function in the barrier region exhibits oscillatory behavior rather than , leading to partial transmission and reflection due to effects at the boundaries. This regime is characterized by analogous to optical waves through a medium with different , where the barrier acts as a relative to the incident energy. The T for a rectangular barrier of width L is given by T = \left[ 1 + \frac{V_0^2 \sin^2(k_2 L)}{4 E (E - V_0)} \right]^{-1}, where k_2 = \sqrt{2m(E - V_0)} / \hbar is the wave number inside the barrier. This expression arises from matching boundary conditions across the three regions, resulting in oscillatory dependence on the phase k_2 L. The \sin^2 term introduces periodic variations in T, with values oscillating between near-zero and unity, resembling Fabry-Pérot interference in . Resonances occur when \sin(k_2 L) \approx 0, or k_2 L = n\pi for integer n, yielding T = 1 and perfect transmission. At these discrete energies, the barrier width accommodates an integer number of half-wavelengths of the particle's de Broglie wave, leading to constructive interference for transmission and zero reflection. The reflection coefficient is R = 1 - T, which vanishes at resonances but is otherwise nonzero, highlighting the non-classical nature of scattering. Even for E > V_0, quantum persists away from resonances due to the abrupt change in wave number at the potential interfaces, causing a mismatch in similar to that at a potential step. This reflection probability decreases with increasing E/V_0 ratio, as the wave numbers k_1 = \sqrt{2mE}/\hbar outside and k_2 inside become more similar. In the high-energy limit where E \gg V_0, T \to 1 and R \to 0, recovering classical free propagation with negligible .

Threshold energy (E = V₀)

At the threshold energy where the incident particle's energy E equals the barrier height V_0, the time-independent inside the barrier (region II, $0 < x < L) reduces to \frac{d^2 \psi}{dx^2} = 0, since the effective wave number \kappa = \sqrt{2m(V_0 - E)}/\hbar = 0. The general solution is thus a linear function, \psi_\mathrm{II}(x) = D + F x, where D and F are constants determined by boundary conditions. This linear form arises directly from integrating the second-order differential equation twice, yielding no curvature in the wave function within the barrier. The transmission coefficient T, obtained by matching the wave function and its derivative at the boundaries x=0 and x=L, is T = \frac{4}{4 + k^2 L^2}, where k = \sqrt{2 m E}/\hbar is the wave number in the free regions (I and III) and L is the full barrier width. Equivalently, T = \left[1 + \left(\frac{k L}{2}\right)^2\right]^{-1}. This expression shows that T decreases monotonically with increasing barrier width L, reflecting the growing phase mismatch across the linear profile; for thick barriers where L \gg 1/k, T approaches zero, indicating near-total reflection. The corresponding reflection coefficient is R = \frac{k^2 L^2}{4 + k^2 L^2}, which satisfies R + T = 1 due to current conservation and approaches unity for wide barriers (L \gg 1/k), consistent with classical expectations at the energy threshold yet modified by quantum boundary effects. This threshold case bridges the sub-barrier regime (exponential evanescence) and the above-barrier regime (interior oscillations), featuring neither exponential suppression nor periodic variation inside the barrier but instead a linear variation that linearly accumulates phase difference over the width L. The absence of decay or oscillation in \psi_\mathrm{II}(x) highlights the transitional nature, where quantum penetration occurs without the probabilistic amplification or damping seen in neighboring energy domains.

Approximations and extensions

WKB approximation for tunneling

The WKB (Wentzel–Kramers–Brillouin) approximation provides a semiclassical method to estimate the transmission probability for quantum tunneling through a potential barrier when the particle energy E is less than the barrier height V_0. In the forbidden region inside the barrier, where E < V(x), the wave function takes the form \psi(x) \sim \exp\left(\pm \int^x \kappa(x') \, dx'\right), with \kappa(x) = \sqrt{2m [V(x) - E]}/\hbar, representing exponentially growing and decaying solutions. Outside the barrier, in the allowed regions, the solutions are oscillatory, \psi(x) \sim \exp\left(\pm i \int^x k(x') \, dx'\right), where k(x) = \sqrt{2m [E - V(x)]}/\hbar. The approximation involves matching these solutions across the classical turning points using connection formulas, which connect the oscillatory behavior to the evanescent waves in the barrier. For a rectangular potential barrier of width a and constant height V_0, the WKB transmission probability simplifies to T \approx 16 \frac{E (V_0 - E)}{V_0^2} \exp\left[-2 \int_0^a \kappa(x) \, dx \right] = 16 \frac{E (V_0 - E)}{V_0^2} \exp(-2 \kappa a), where \kappa = \sqrt{2m (V_0 - E)} / \hbar is constant. This expression captures the dominant exponential suppression of tunneling, with the prefactor accounting for the amplitude matching at the boundaries. The integral \int_0^a \kappa(x) \, dx represents the action in the forbidden region, generalizing to arbitrary smooth potentials as \exp\left[-2 \int_{x_1}^{x_2} \kappa(x) \, dx \right], where x_1 and x_2 are the turning points. The WKB approximation is valid for potentials that vary slowly compared to the local de Broglie wavelength, specifically when \left| \frac{d\lambda}{dx} \right| \ll 1, or equivalently \left| \frac{dV}{dx} \right| \ll \frac{p^2}{m \lambda}, ensuring the wave function changes gradually. For rectangular barriers, which have abrupt changes at the edges, the approximation incurs errors near the turning points but remains accurate overall for thick barriers where \kappa a \gg 1, as the edge contributions become negligible. In this regime, it excels for smooth or gently varying potentials but provides a reliable estimate even for rectangular cases when the barrier is sufficiently wide. Compared to the exact solution for the , the WKB form matches the exponential decay factor by approximating \sinh(\kappa a) \approx \frac{1}{2} \exp(\kappa a) for large \kappa a, which dominates the transmission probability expression T = \left[1 + \frac{V_0^2 \sinh^2(\kappa a)}{4 E (V_0 - E)}\right]^{-1}. This yields the correct leading exponential behavior while the prefactor $16 E (V_0 - E)/V_0^2 emerges from the boundary matching, providing an exact match to the exact prefactor in the high-opacity limit. Thus, for \kappa a \gg 1, the WKB result is precise up to the prefactor, making it a powerful tool for estimating tunneling in thick barriers.

Ramsauer-Townsend effect

The Ramsauer-Townsend effect describes the pronounced minimum in the total elastic scattering cross-section observed when low-energy electrons interact with noble gas atoms, such as , , and . This minimum typically occurs at electron kinetic energies on the order of 0.1 to 1 eV, where the cross-section drops to nearly zero, indicating almost perfect transmission without scattering. The phenomenon was first experimentally documented in 1921 by through measurements of electron transmission through noble gases, revealing an unexpected dip in scattering efficiency at these energies. Independently, reported similar findings in the same year using beam attenuation techniques on rarefied gases. This effect underscores the quantum mechanical nature of low-energy electron-atom interactions, contrasting with classical expectations of monotonic increase in scattering at low velocities. In the rectangular potential barrier model, the short-range attractive interaction between the electron and the noble gas atom is approximated by a one-dimensional rectangular potential well of depth V_0 > 0 and width $2a, with V_0 chosen to mimic the effective potential's scale (typically a few ). This simple model captures the essential responsible for the effect, treating the atom as a of lower potential over scales (a \approx 1 ). The probability T through the well oscillates with energy due to multiple reflections at the boundaries, analogous to optical Fabry-Pérot resonances. In the low-energy regime, the wave function is oscillatory both outside and inside the well, with T reaching unity at resonances. The minimum cross-section arises from a phase shift in the scattered wave that causes destructive interference in the outgoing waves. In the one-dimensional model, this manifests as resonant transmission where T = 1, effectively eliminating backscattering. The condition for resonance occurs when the well width aligns with the de Broglie wavelength \lambda of the electron inside the well region, specifically satisfying $2 k_2 a = n \pi, where k_2 = \sqrt{2m (E + V_0)} / \hbar, m is the electron mass, and n is a positive integer; the first resonance (n=1) appears at low E \approx \frac{\pi^2 \hbar^2}{8 m a^2} - V_0. This geometric matching implies $2a \approx n \lambda_2 / 2, where \lambda_2 = 2\pi / k_2 is the wavelength inside the well. In three dimensions, this translates to an s-wave phase shift \delta_0 = n \pi, yielding zero cross-section \sigma \propto (4\pi / k^2) \sin^2 \delta_0 = 0 at the Ramsauer energy. For E \to 0, while T \approx [4 k a / (V_0 / \hbar v)]^2 in perturbative limits (with k = \sqrt{2m E}/\hbar and v the velocity), the dominant feature is the approach to unit transmission at the phase condition, enabling the effective zero cross-section. Experimentally, the effect highlights the inertness of , as their filled electron shells create a symmetric potential leading to near-transparent at Ramsauer energies, reducing collision rates in low-pressure gases. Modern calculations using the rectangular model reproduce observed minima (e.g., ~0.7 eV for , ~1.2 eV for ) by tuning V_0 to atomic ionization potentials and a to van der Waals radii, validating its utility despite simplifications.

Applications

Quantum tunneling phenomena

Quantum tunneling through a potential barrier lower than the particle's energy is a cornerstone of quantum mechanics, but the rectangular potential barrier model particularly illuminates phenomena where particles with energies below the barrier height (E < V₀) escape via tunneling. One of the earliest and most impactful applications is in alpha decay, where atomic nuclei emit helium-4 nuclei (alpha particles). In 1928, George Gamow developed a quantum mechanical model explaining alpha decay as the tunneling of a preformed alpha particle through the Coulomb barrier surrounding the nucleus. Although the actual barrier is Coulombic (V(r) ∝ 1/r), Gamow's approach approximated it as rectangular for analytical simplicity in early calculations, treating the barrier as a region of constant height V₀ over a finite width. This simplification captures the essence of the tunneling probability T ≈ exp[-2 ∫ κ(r) dr], where κ(r) = √[2m(V(r) - E)] / ℏ; for a rectangular barrier with constant V, the integral reduces to κ L, yielding T ∝ exp(-2 κ L) with κ = √[2m(V₀ - E)] / ℏ. The decay rate Γ, proportional to the tunneling probability, follows Γ ∝ exp[-2 ∫ κ(r) dr], directly linking to barrier penetration. While the rectangular approximation provided initial insights, the quantitative prediction of alpha decay half-lives that matched the empirical Geiger-Nuttall law—an relation between range (related to energy) and decay constant observed in 1911—arises from Gamow's full model using the . This law, log(λ) ∝ 1/√E_α (where λ is the decay constant and E_α the energy), emerges from the exponential dependence involving ∫ dr / √(V(r) - E) over the 1/r potential, validating the tunneling mechanism and resolving the classical puzzle of why , bound by high barriers (~25 MeV), escape in measurable times (from microseconds to billions of years). Gamow's work not only confirmed quantum tunneling's role in nuclear stability but also influenced the liquid-drop model of the . Another key phenomenon is field emission, or cold electron emission from metals under strong electric fields, where electrons tunnel from the Fermi level through the surface potential barrier at low temperatures (near 0 K). In 1928, Ralph Fowler and Lothar Nordheim formulated the theory using a triangular barrier shaped by the applied field, but the rectangular barrier serves as a useful approximation for thin insulating films or narrow barriers where the potential is nearly constant. The emission current I is proportional to the transmission coefficient T at the Fermi energy E_F, with T ≈ exp[-2 ∫ κ(x) dx]; for a rectangular barrier of height φ (work function) and width w ≈ φ / eF, this approximates to I ∝ T(E_F) ∝ exp\left[ -\frac{2 \sqrt{2m} \phi^{3/2}}{\hbar e F} \right], providing an order-of-magnitude estimate for the exponential field dependence, though the more precise triangular barrier yields exp\left[ -\frac{8\pi \sqrt{2m} \phi^{3/2}}{3 \hbar e F} \right]. In cold emission scenarios, the barrier width decreases with increasing voltage, enhancing tunneling without thermal activation, as electrons occupy states up to E_F and tunnel directly. This rectangular model approximates scenarios in thin-film devices, where the barrier is effectively flat over short distances, providing a baseline for understanding field-lowered barriers in vacuum electronics.

Device physics examples

The tunnel diode, also known as the Esaki diode, was invented in 1957 by while studying heavily doped p-n junctions at . This device exploits resonant tunneling through a potential barrier formed at the p-n junction, where electrons with energies matching the valence band maximum on the p-side can tunnel to the conduction band minimum on the n-side, resulting in a current peak followed by a valley in the I-V characteristic that exhibits negative differential resistance. The negative resistance region enables applications in high-frequency oscillators and amplifiers, with the barrier often approximated as a single rectangular potential for basic models, though double rectangular barriers are used to capture resonant peaks in more detailed simulations of the transport mechanism. In non-volatile flash memory devices, such as and flash, Fowler-Nordheim tunneling through a thin insulating layer serves as the primary for programming and erasing charge on a floating gate. The tunnel , typically ≈5 nm thick (SiO₂) with an electron barrier height V₀ ≈ 3.1 eV relative to , allows electrons to tunnel under high (E < V₀) when a voltage is applied, enabling charge injection or removal to store binary states. This process, modeled using the rectangular barrier approximation for the potential, achieves non-volatile retention by confining charges against thermal escape, though thickness below 5 nm increases direct tunneling leakage, limiting to around 10⁵–10⁶ cycles in modern cells. Scanning tunneling microscopy (), invented in 1981 by and , relies on quantum tunneling of electrons between a sharp metallic tip and a sample surface separated by a gap of ~0.5–1 nm, modeled as a rectangular potential barrier of height ~4–5 eV ( difference). The tunneling current, exponentially sensitive to gap distance d via I ∝ exp(-2κ d) with κ ≈ √(2m φ)/ℏ, allows atomic-resolution imaging by maintaining while scanning the tip, with the rectangular model providing the basis for interpreting current-distance characteristics in simple theories. Semiconductor quantum dots and quantum wells employ finite rectangular potential barriers to confine electrons in zero- or two-dimensional structures, such as InAs dots in GaAs matrices, where the barrier height (typically 0.5–1 ) defines discrete energy levels for optoelectronic applications like lasers and detectors. However, the finite barrier nature permits evanescent wave penetration and tunneling leakage of carriers out of the confined region, which degrades device performance by reducing and efficiency; for instance, in quantum dot LEDs, this leakage contributes to non-radiative recombination, shortening operational lifetime to hours under high injection unless mitigated by barrier engineering. Modern device extensions approximate rectangular barriers in novel materials for enhanced tunneling control. In graphene-based nanostructures, electrostatic gating creates sharp rectangular potential barriers (widths ~10–100 nm, heights ~0.1–1 ) that enable near-perfect Klein tunneling transmission at normal incidence due to the Dirac-like dispersion, facilitating applications in high-mobility transistors and valley filters. Similarly, in superconducting Josephson junctions, thin insulating barriers (e.g., AlOₓ, ~1–2 nm thick) are modeled as rectangular potentials for tunneling, supporting dissipationless supercurrents up to a critical I_c ≈ (πΔ/2eR_N) tanh(Δ/2kT), where Δ is the superconducting and R_N the normal , underpinning qubits and sensitive magnetometers in .

References

  1. [1]
    [PDF] Quantum Mechanics - Lehman College
    Mar 5, 2019 · Rectangular Potential Barrier - Tedious Derivations where , , and are the incident, reflected, and transmitted probability currents ...
  2. [2]
    [PDF] Lecture 2 Quantum mechanics in one dimension - TCM
    For quantum particles incident on a spatially localized potential barrier, the scattering properties are defined by a unitary S-matrix, ψout = Sψin . The ...
  3. [3]
    [PDF] chapter 4
    Plot barrier width versus transmission coefficient for a rectangular barrier of width 2a for the case E = V0. What widths of the barrier result in 10%, 1 ...<|control11|><|separator|>
  4. [4]
    The Quantum Tunneling of Particles through Potential Barriers
    Quantum tunneling is a phenomenon in which particles penetrate a potential energy barrier with a height greater than the total energy of the particles.
  5. [5]
  6. [6]
    7.6 The Quantum Tunneling of Particles through Potential Barriers
    Sep 29, 2016 · A year later, George Gamow used the formalism of quantum mechanics to explain the radioactive α α -decay of atomic nuclei as a quantum-tunneling ...Missing: history | Show results with:history
  7. [7]
  8. [8]
    [PDF] Notes on Quantum Mechanics - Physics
    Figure 3.12: A rectangular potential barrier (a) with V = 0.05 inside, and scattering solutions of the TISE for this potential with E = 0.04, so that the ...<|control11|><|separator|>
  9. [9]
    [PDF] PHYSICS 430 Lecture Notes on Quantum Mechanics
    ... Rectangular Potentials. Qualitative behavior of energy eigenstates. Why bound state energies are dis- crete, why unbound eigenstates oscillate. Step potentials.<|control11|><|separator|>
  10. [10]
    None
    Below is a merged summary of the Rectangular Potential Barrier from Chapter 2, Section 2.5, around Problem 2.32, as presented in Griffiths' *Introduction to Quantum Mechanics* (3rd ed, 2018). The information from all provided summaries has been consolidated into a detailed response, with a table in CSV format to organize the key details efficiently. The response retains all unique information, resolves redundancies, and provides a comprehensive overview.
  11. [11]
    [PDF] Physics 386
    The positive exponential corresponds to right moving particles and the negative to left moving particles. We are free to use just a single constant, B, because ...
  12. [12]
    cover page
    ### Transmitted Wave in Region III for Rectangular Potential Barrier
  13. [13]
    [PDF] A Review of Quantum Mechanics - Rutgers Physics
    FIG. 2: Wave numbers for a rectangular potential barrier. Its probability current density is conserved across the boundary: j< ...<|control11|><|separator|>
  14. [14]
    None
    ### Extracted Expressions from Griffiths QM 3e, Problem 2.33
  15. [15]
    [PDF] Quantum Mechanics in One Dimension, Part II - Texas Tech University
    Sep 17, 2020 · Transmission Coefficient for the Rectangular Barrier. Note that cosix = chx, sinix = i shx, ch2x − sh2x = 1. The reflection coefficient is. R ...
  16. [16]
  17. [17]
    [PDF] 4.4 The Potential Step - Hansraj College
    If E w V0, and hence w 1, the transmission coefficient T becomes asymptotically equal to unity, T. 1, and R. 0. So, at very high energies and weak potential ...
  18. [18]
    WKB Approximation - Richard Fitzpatrick
    The WKB approximation only applies to situations in which there is very little chance of a particle tunneling through the potential barrier in question.
  19. [19]
    [PDF] Quantum Physics III Chapter 3: Semiclassical Approximation
    Feb 3, 2019 · This is a superposition of two waves, a wave ψinc incident on the barrier and a wave ψref reflected from the barrier. The incident part is.
  20. [20]
    (PDF) Resonant transmission though rectangular potential well
    Oct 4, 2024 · The approach provides satisfactory explanation of the Ramsauer-Townsend effect and it is within the reach of the audience at the undergraduate ...
  21. [21]
    [PDF] Zur Quantentheorie des Atomkernes
    Gamow~ z. Zt. in GSttingen. Mit 5 Abbildungen. (Eingegangen am 2. August 1928.) Es wird der Versuch gemacht, ...Missing: Aufbaues | Show results with:Aufbaues
  22. [22]
    [PDF] Ranges of a particles from Radioactive Substances. 613
    Geiger and Mr. J. M. Nuttall on the Ranges. --in the case of uranlum--have recently been giw,,n by. Geiger and Rutherford*, and by Focht. It would be of great ...
  23. [23]
    Electron emission in intense electric fields - Journals
    The main features of the phenomenon of the extraction of electrons from cold metals by intense electric fields are well known.
  24. [24]
    [PDF] University of Virginia
    a rectangular potential barrier for the insulating gap and using the electron free mass in both the metals and the insulator. Sommerfeld and Bethe [2]- were ...
  25. [25]
    Review on Non-Volatile Memory with High-k Dielectrics - NIH
    Since the silicon-oxide barrier height is 3.2 eV, the drain voltage has to be at least greater than 3.2 V for reasonable efficiency. Therefore, there is a major ...