Fact-checked by Grok 2 weeks ago

Step potential

In , the step potential is an idealized one-dimensional model of a function defined as V(x) = 0 for x < 0 and V(x) = V_0 (a constant) for x \geq 0, used to analyze the scattering behavior of particles incident on a sudden change in potential. This setup illustrates key quantum phenomena such as wave reflection, transmission, and penetration into classically forbidden regions, contrasting with classical predictions where particles with energy E < V_0 would be fully reflected without entering the barrier. For incident particles with energy E > V_0, the wave function consists of an incoming plane wave and a reflected wave in the region x < 0, and a transmitted wave in x > 0, leading to partial reflection and transmission. The reflection coefficient R = \left| \frac{k - k'}{k + k'} \right|^2 and transmission coefficient T = \frac{4 k k'}{(k + k')^2}, where k = \sqrt{2mE}/\hbar and k' = \sqrt{2m(E - V_0)}/\hbar, satisfy R + T = 1, demonstrating conservation of probability current despite the abrupt potential change. In the case E < V_0, classical mechanics predicts total reflection, but quantum mechanically, the wave function decays exponentially in the forbidden region x > 0 as \psi(x) \propto e^{-\kappa x} with \kappa = \sqrt{2m(V_0 - E)}/\hbar, allowing a nonzero probability density to penetrate the barrier, though R = 1 and T = 0. This model serves as a foundational example for understanding more complex potentials, such as barriers and wells, and highlights the wave-like nature of particles in scattering theory.

Definition and Setup

Potential Profile

The step potential is a fundamental model in one-dimensional , characterized by a constant function V(x) that jumps discontinuously at x = 0. Specifically, it is defined as V(x) = 0 for x < 0 and V(x) = V_0 for x \geq 0, where V_0 > 0 represents the height of the step. This setup creates a sharp boundary between two regions of differing , with the left side (x < 0) acting as a free-particle region and the right side (x \geq 0) as a higher-energy barrier. Graphically, the potential profile resembles a vertical step or infinite barrier originating at x = 0, extending indefinitely to the right with constant height V_0, while remaining at zero to the left. This idealized form simplifies the analysis of quantum scattering while capturing the essence of discontinuous potential changes. Physically, the step potential serves as a model for abrupt transitions in potential energy, such as those encountered in semiconductor heterostructures where materials with different band gaps are joined, or in idealized problems of particle scattering off barriers. In typical setups, a particle with total energy E is assumed to be incident from the left region (x < 0), allowing exploration of quantum effects like reflection and transmission at the interface. This configuration plays a key role in solving the time-independent Schrödinger equation for scattering states.

Time-Independent Schrödinger Equation

The time-independent governs the spatial behavior of stationary quantum states in one dimension, providing an eigenvalue equation for the wave function \psi(x) and energy E. Originally derived by in his seminal 1926 paper, it takes the form -\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, and V(x) is the potential energy function. This equation arises from separating variables in the full time-dependent for time-independent potentials, yielding solutions of the form \psi(x) e^{-iEt/\hbar} that describe stationary states with definite energy. For the step potential, defined previously as V(x) = 0 for x < 0 and V(x) = V_0 for x > 0, the equation applies to reflect the regions of constant potential. In the region x < 0, where the particle behaves as a free particle, the equation simplifies to -\frac{\hbar^2}{2m} \frac{d^2 \psi(x)}{dx^2} = E \psi(x), with solutions characterized by the wave number k_1 = \sqrt{2mE}/\hbar. For x > 0, the constant potential shifts the effective energy, yielding -\frac{\hbar^2}{2m} \frac{d^2 \psi(x)}{dx^2} + V_0 \psi(x) = E \psi(x), and the corresponding wave number k_2 = \sqrt{2m(E - V_0)}/\hbar, which is real when E > V_0. Here, E represents the total energy of the particle, serving as the eigenvalue in this second-order differential equation. This setup frames the step potential as an eigenvalue problem, where solutions correspond to scattering states relevant to quantum transmission and reflection, always emphasizing the stationary nature of the wave functions. The constants \hbar \approx 1.0545718 \times 10^{-34} J s and m (typically on the order of electron mass $9.109 \times 10^{-31} kg for common applications) ensure dimensional consistency, with energy in joules and length in meters.

Solution for the Step Potential

General Wave Function Forms

In the step potential problem, the time-independent is solved separately in the two regions defined by the potential: region I for x < 0 where V(x) = 0, and region II for x > 0 where V(x) = V_0. The general solutions take the form of plane waves or exponential functions depending on the particle's energy E relative to V_0. These solutions assume a time-dependent of the form \psi(x, t) = \psi(x) e^{-i E t / \hbar}, where \psi(x) is the spatial part, and the waves are normalized such that the incident flux is unity. For E > 0 in region I (x < 0), the general wave function consists of an incident wave traveling to the right and a reflected wave traveling to the left: \psi_1(x) = A e^{i k_1 x} + B e^{-i k_1 x}, where A and B are complex coefficients representing the amplitudes of the incident and reflected waves, respectively, and k_1 is the wave number in this region. In region II (x > 0), the form of the wave function depends on whether E > V_0 or E < V_0. For E > V_0, the solution is a transmitted plane wave propagating to the right: \psi_2(x) = C e^{i k_2 x}, where C is the transmission amplitude and k_2 is the wave number in this region; there is no leftward-propagating wave due to the absence of reflection from infinity. For E < V_0, the solution is an evanescent wave that decays exponentially: \psi_2(x) = C e^{-\kappa x}, where \kappa = \sqrt{2m(V_0 - E)} / \hbar characterizes the decay rate, with m the particle mass and \hbar the reduced Planck's constant; this form ensures the wave function remains finite as x \to \infty. The wave numbers k_1 and k_2 are determined from the Schrödinger equation in their respective regions.

Matching Boundary Conditions

To determine the coefficients relating the incident, reflected, and transmitted waves in the step potential, the continuity requirements of the wave function and its first derivative at the boundary x = 0 are imposed. These conditions arise because the time-independent Schrödinger equation is a second-order differential equation, requiring that \psi(x) and \psi'(x) remain continuous across the finite potential step to ensure a physically acceptable, single-valued, and finite solution everywhere. Assuming the general wave function forms from the solutions in each region—with \psi_I(x) = A e^{i k_1 x} + B e^{-i k_1 x} for x < 0 (where k_1 = \sqrt{2mE}/\hbar) and \psi_{II}(x) accordingly for x > 0—the boundary matching proceeds separately for the cases E > V_0 and E < V_0. For E > V_0, the transmitted wave is oscillatory, so \psi_{II}(x) = C e^{i k_2 x} with k_2 = \sqrt{2m(E - V_0)}/\hbar. Applying continuity at x = 0: \psi_I(0) = \psi_{II}(0) \implies A + B = C \psi_I'(0) = \psi_{II}'(0) \implies i k_1 (A - B) = i k_2 C Simplifying the derivative equation gives k_1 (A - B) = k_2 C. Solving these simultaneously yields the reflection amplitude B/A = (k_1 - k_2)/(k_1 + k_2) and the transmission amplitude C/A = 2 k_1 / (k_1 + k_2). For E < V_0, the region x > 0 is classically forbidden, and the wave function is evanescent: \psi_{II}(x) = C e^{-\kappa x} with \kappa = \sqrt{2m(V_0 - E)}/\hbar > 0. The continuity conditions at x = 0 are: \psi_I(0) = \psi_{II}(0) \implies A + B = C \psi_I'(0) = \psi_{II}'(0) \implies i k_1 (A - B) = -\kappa C Solving these equations results in the reflection amplitude B/A = (k_1 - i \kappa)/(k_1 + i \kappa) and the transmission amplitude C/A = 2 k_1 / (k_1 + i \kappa). These relations are unique to the scattering configuration, where an incident wave approaches from the left (x < 0) and there is no incoming wave from the right (x > 0). This setup justifies omitting the e^{-i k_2 x} term for E > V_0 (which would represent an incident wave from +\infty) and the divergent e^{\kappa x} term for E < V_0 (to ensure the wave function remains normalizable or bounded at large positive x).

Reflection and Transmission

Coefficients for E > V₀

When the particle energy E exceeds the step height V_0, both regions support propagating , leading to partial and . The ratios of the amplitudes B/A and C/A are determined by matching the wave function and its derivative at the boundary x=0, yielding B/A = (k_1 - k_2)/(k_1 + k_2) and C/A = 2k_1/(k_1 + k_2), where k_1 = \sqrt{2mE}/\hbar and k_2 = \sqrt{2m(E - V_0)}/\hbar. The R, representing the fraction of the incident that is reflected, is given by R = \left| \frac{B}{A} \right|^2 = \left( \frac{k_1 - k_2}{k_1 + k_2} \right)^2. This expression arises directly from the and the identical numbers in the incident and reflected regions. The T accounts for the difference in speeds across the step and is defined using to ensure . The incident is j_\text{inc} = (\hbar k_1 / m) |A|^2, while the transmitted is j_\text{trans} = (\hbar k_2 / m) |C|^2. Thus, T = \frac{j_\text{trans}}{j_\text{inc}} = \frac{k_2}{k_1} \left| \frac{C}{A} \right|^2 = \frac{4 k_1 k_2}{(k_1 + k_2)^2}. requires R + T = 1, which holds upon substitution of the . As E increases, k_2 approaches k_1, causing R to decrease toward zero and T to approach unity, consistent with classical expectations for high energies where the step becomes negligible.

Behavior for E < V₀

When the energy E of the incident particle is less than the step height V_0, the region x > 0 is classically forbidden, yet predicts total with into this . The wave in the forbidden takes an evanescent form, obtained by matching boundary conditions at x = 0, consisting of a decaying rather than an oscillating wave. The is R = |B/A|^2 = 1, indicating complete reflection of the incident back to the left. The T = 0, even though the amplitude C/A in the forbidden region is non-zero, because the evanescent wave carries no net probability flux due to the imaginary wave number k_2 = i\kappa, where \kappa = \sqrt{2m(V_0 - E)} / \hbar > 0. The reflected wave acquires a phase shift given by \arg(B/A) = -2 \tan^{-1}(\kappa / k_1), where k_1 = \sqrt{2mE} / \hbar, resulting in an energy-dependent difference relative to the incident wave. This shift arises from the reflection amplitude B/A = (k_1 - i\kappa)/(k_1 + i\kappa). In the forbidden region, the probability density |\psi_2(x)| decays exponentially as e^{-2\kappa x}, characterizing the penetration depth as $1/\kappa, which decreases as E approaches V_0 from below.

Interpretation of Results

Probability Currents

In quantum mechanics, the probability current in one dimension provides a measure of the flow of probability density associated with a particle's wave function. For a wave function \psi(x), the probability current j(x) is defined as j(x) = \frac{\hbar}{2mi} \left[ \psi^*(x) \frac{d\psi(x)}{dx} - \psi(x) \frac{d\psi^*(x)}{dx} \right], where \hbar is the reduced Planck's constant and m is the particle mass. This expression arises from the continuity equation \frac{\partial |\psi|^2}{\partial t} + \frac{\partial j}{\partial x} = 0, ensuring local conservation of probability. For the step potential, where the wave function in the region x < 0 takes the form \psi(x) = A e^{i k_1 x} + B e^{-i k_1 x} with k_1 = \sqrt{2mE}/\hbar and in x > 0 as \psi(x) = C e^{i k_2 x} with k_2 = \sqrt{2m(E - V_0)}/\hbar for E > V_0, the currents simplify due to the plane-wave components. The incident current from the right-moving wave is j_\text{inc} = \frac{\hbar k_1}{m} |A|^2 > 0, representing the incoming probability flux. The reflected current from the left-moving wave is j_\text{ref} = -\frac{\hbar k_1}{m} |B|^2 < 0, indicating flux directed back toward negative x. The transmitted current is j_\text{trans} = \frac{\hbar k_2}{m} |C|^2 > 0. Conservation of probability requires that the net current be uniform across the potential step in the , yielding j_\text{inc} + j_\text{ref} = j_\text{trans}. This relation validates the boundary-matching conditions by confirming that the incident flux equals the outgoing reflected and transmitted fluxes in magnitude, with the R = |B/A|^2 and T = (k_2 / k_1) |C/A|^2 satisfying R + T = 1. For E < V_0, the transmitted wave is evanescent (k_2 \to i \kappa with \kappa > 0), resulting in j_\text{trans} = 0, total reflection (R = 1), and j_\text{inc} + j_\text{ref} = 0. The underlying guarantee of this current conservation stems from the Hermitian nature of the operator \hat{H} in the time-independent . A Hermitian \hat{H} ensures that the preserves the of the wave function, as the rate of change of total probability vanishes: \frac{d}{dt} \int |\psi|^2 dx = 0. For the step potential, where \hat{H} = -\frac{\hbar^2}{2m} \frac{d^2}{dx^2} + V(x) with real V(x), hermiticity holds, enforcing steady-state current continuity.

Physical Implications

In , the step potential reveals a counterintuitive phenomenon where a particle with E > V_0 experiences partial despite having sufficient to classically surmount the step, arising from an impedance mismatch between the differing wave numbers k_1 = \sqrt{2mE}/\hbar in region I and k_2 = \sqrt{2m(E - V_0)}/\hbar in region II, which enforces continuity of the wave and its at the . This quantum , with a probability R = |(k_1 - k_2)/(k_1 + k_2)|^2 > 0, contrasts sharply with classical expectations of full transmission, highlighting the wave-like interference effects inherent to the . For E < V_0, the step potential leads to total reflection (R = 1) accompanied by a phase shift in the reflected wave, as the wave function penetrates evanescently into region II without propagating, underscoring the particle's wave nature in a regime devoid of a direct classical counterpart beyond simple rebound. This behavior manifests the quantum prohibition on particles entering classically forbidden regions while allowing a decaying tail of the probability density, which has no classical analogue in terms of oscillatory penetration. The WKB (Wentzel-Kramers-Brillouin) approximation, which assumes slowly varying potentials to semiclassically estimate wave functions, performs poorly near the abrupt step due to the infinite force (delta-function-like derivative) at the discontinuity, violating the condition that the potential change occurs over distances much larger than the . In contrast, the approximation fares better for smooth potentials where the change is gradual, allowing more accurate predictions of transmission and reflection away from sharp turning points. The reflection R(E) and transmission T(E) coefficients derived from the step potential are directly measurable in quantum scattering experiments, such as those involving low-energy electron beams or atomic interferometry, providing empirical validation of wave mechanics. This setup bears analogy to the observed in noble gas scattering, where energy-dependent minima in cross sections reflect resonant-like transmission behaviors akin to those in step-like potentials.

Extensions and Variations

Negative Step Potential

In the negative step potential, the potential energy function is defined as V(x) = 0 for x < 0 and V(x) = V_0 for x > 0, where V_0 < 0. This configuration represents a downward step, allowing the incident particle with energy E > 0 to enter a of lower potential, which classically would result in full without . The wave functions in both regions are oscillatory, with the transmitted wave exhibiting higher momentum due to the reduced potential. Specifically, the wave number in the left region is k_1 = \sqrt{2mE}/\hbar, while in the right region it is k_2 = \sqrt{2m(E - V_0)}/\hbar, satisfying k_2 > k_1 since V_0 < 0. Unlike the positive step case, there is no evanescent wave because E > V_0 holds everywhere for E > 0, ensuring propagating solutions on both sides. The reflection and transmission coefficients retain the same functional form as in the positive step potential but yield distinct behavior due to k_2 > k_1. The is R = \left[ (k_1 - k_2)/(k_1 + k_2) \right]^2, which is smaller than in the positive step case and decreases as |V_0| increases. The is T = 4 k_1 k_2 / (k_1 + k_2)^2, approaching unity more rapidly with higher energy or deeper step, satisfying R + T = 1 from conservation. For example, if E = |V_0|/3, then R = 1/9 and T = 8/9. Physically, this setup demonstrates quantum reflection arising from the abrupt change in wave speed, analogous to an impedance mismatch, despite the classical expectation of in the lower potential region without backscattering. The higher k_2 implies increased post-transmission, highlighting wave-particle duality in potential discontinuities.

Relativistic Formulation

The relativistic formulation of the step potential employs the to describe the behavior of relativistic electrons encountering a potential discontinuity. The is expressed as i \hbar \frac{\partial \psi}{\partial t} = \left[ c \vec{\alpha} \cdot \vec{p} + \beta m c^2 + V(x) \right] \psi, where \psi is a four-component , \vec{\alpha} and \beta are Dirac matrices, V(x) represents the (typically V(x) = 0 for x < 0 and V(x) = V_0 for x > 0), acting as an electrostatic coupled to the charge. Solutions to this equation consist of plane waves modulated by . For an incident from the left (x < 0), the wave function includes an incident component, a reflected component, and, for x > 0, a transmitted component, all constructed from positive-energy spinors for E > m c^2. At the boundary x = [0](/page/0), continuity of the full four-component spinor \psi is imposed to match the wave functions across the step. In the non-relativistic limit where particle velocity v \ll c, these solutions reduce to the familiar results for the step potential. A striking feature emerges in the so-called Klein zone, where the incident energy satisfies m c^2 < E < V_0 - m c^2 (assuming V_0 > 2 m c^2). Here, the transmission coefficient satisfies |T| > 1, while the reflection coefficient |R| > 0, violating unitarity in single-particle interpretation. This Klein paradox arises because the transmitted wave corresponds to a negative-energy state in the potential region, interpreted in quantum field theory as pair creation: the incident electron promotes a valence electron from the Dirac sea, producing a positron that propagates leftward (appearing as enhanced reflection) and a transmitted hole. For a negative step potential (V_0 < 0), the transmission is enhanced compared to the non-relativistic case, approaching unity for relativistic incident particles without invoking the , as the potential drop facilitates propagation without pair production. In the Klein-Gordon equation for spin-0 particles with a scalar potential (coupling to mass), a negative step further reduces the effective mass in the x > 0 region, leading to similarly enhanced transmission probabilities while avoiding the negative probability currents associated with the in vector potentials.

Comparisons and Limits

Classical Analogue

In classical mechanics, the analogue of the quantum step potential V(x) = 0 for x < 0 and V(x) = V_0 for x > 0 describes the trajectory of a particle incident from the left (x < 0) with total energy E. For E > V_0, the particle encounters no barrier to entry into the region x > 0; it crosses the step at x = 0 without reflection, continuing to the right indefinitely. Upon crossing, the particle's kinetic energy decreases from E to E - V_0 due to conservation of total energy, resulting in a velocity change from v_1 = \sqrt{\frac{2E}{m}} before the step to v_2 = \sqrt{\frac{2(E - V_0)}{m}} after the step, where m is the particle's mass. This slowdown reflects the abrupt increase in potential, but the motion remains deterministic and unidirectional. For E < V_0, the region x > 0 is classically inaccessible, as the particle lacks sufficient to overcome the step. The particle approaches from the left with v = \sqrt{\frac{2E}{m}}, reaches the step at x = 0 (the classical ), and undergoes total elastic reflection, reversing direction without energy loss. The velocity magnitude remains \sqrt{\frac{2E}{m}}, but its sign changes from positive to negative, conserving both and magnitude while flipping the momentum direction in . Unlike , where and probabilities are partial even for E > V_0, the classical analogue yields deterministic outcomes: full (T = 1, R = 0) for E > V_0 and full (T = 0, R = 1) for E < V_0, with no probabilistic interpretation required.

Finite Barrier Connection

The step potential can be understood as a limiting case of the finite rectangular potential barrier, where the barrier width L approaches infinity while maintaining a fixed height V_0. In this limit, for incident particle energies E < V_0, the transmission probability T through the finite barrier vanishes exponentially, approaching zero, which aligns precisely with the complete observed in the step potential, where no transmitted wave exists beyond the . Conversely, for E > V_0, the transmission and reflection coefficients of the finite barrier converge to those of the step potential, yielding a non-zero T that depends on the mismatch between the wave numbers on either side of the . For a finite barrier width L, the transmission probability T(E) exhibits oscillatory behavior as a function of energy, featuring sharp resonances where T approaches unity at specific energies corresponding to quasi-bound states within the barrier. These resonances arise from interference between waves reflected at the two boundaries of the barrier, enabling perfect transmission under certain conditions. In contrast, the step potential lacks such resonances because it possesses only a single interface, preventing the formation of standing waves or multiple reflections that characterize the finite case. In the region where E < V_0, both the step potential and the finite barrier involve evanescent waves that decay exponentially away from the interface, illustrating the penetration of the wave function into the classically forbidden region. For the step potential, this decay extends indefinitely without transmission, embodying an infinite barrier scenario. The finite barrier, however, permits a small but non-zero transmission through evanescent coupling across the width L, suppressed exponentially by factors involving \sqrt{2m(V_0 - E)} L / \hbar, which diminishes as L increases. Mathematically, the connection between the two potentials stems from the shared use of boundary condition matching at potential discontinuities to ensure continuity of the wave function and its derivative. The step potential simplifies this to a single interface, whereas the finite barrier requires matching at two interfaces, leading to more complex coefficient relations but reducing to the step form in the infinite-width limit. This continuity underscores the step potential's role as a foundational model for understanding scattering in piecewise constant potentials.

Applications in Physics

Quantum Tunneling Contexts

In the step potential scenario where the particle energy E is less than the potential height V_0, the quantum wave function \psi(x) extends into the classically forbidden region for x > 0, decaying exponentially and yielding a non-zero there, in stark contrast to classical expectations of zero penetration. Despite this extension, the R = 1, ensuring total reflection with no transmitted wave, and the net in the forbidden region is zero, precluding any true transmission or tunneling to . This behavior highlights a key quantum feature: probabilistic presence without flux, distinguishing it from classical impenetrability. The step potential approximates situations with abrupt potential rises, such as in field emission where electrons encounter a sudden rise at a metal-vacuum , modeling the initial barrier though the model's infinite width restricts full transmission calculations compared to realistic finite barriers. Similarly, it illustrates the phase in nuclear processes like , where the alpha particle faces a steep , but the infinite extent limits direct quantitative application to decay rates. The evanescent decay length $1/\kappa, with \kappa = \sqrt{2m(V_0 - E)}/\hbar, quantifies this . Stationary-state solutions suggest perpetual wave function overlap in the forbidden region, implying eternal "tunneling"; however, time-dependent analyses using localized wave packets, such as Gaussians, reveal transient penetration, where the packet's tail briefly enters x > 0 before reflecting, allowing definition of a finite tunneling time via phase delays or dwell times. Experimental realizations include neutron optics experiments, where low-energy neutrons totally reflect at refractive index steps due to nuclear potentials at material interfaces, mimicking the quantum step reflection. Electron diffraction setups with sharp potential gradients in crystals also demonstrate these effects through observed reflection patterns.

Modern Quantum Devices

In semiconductor heterojunctions, such as those formed between GaAs and AlGaAs, the abrupt conduction band offset is modeled as a step potential in the effective mass approximation, which confines electrons to form a (2DEG) at the interface. This step-like potential arises from the difference in band alignments, typically around 0.3 eV for AlGaAs/GaAs, leading to quantum confinement in the growth direction and high-mobility transport in the plane. Seminal experiments in the 1970s demonstrated 2DEG formation with mobilities exceeding 10^5 cm²/Vs at low temperatures, enabling applications in high-electron-mobility transistors (HEMTs). Quantum cascade lasers (QCLs) employ sequential step potentials within coupled quantum wells to facilitate intersubband transitions, where electrons cascade through multiple stages, emitting photons at mid-infrared wavelengths. The active region consists of injector and transition regions modeled as finite step barriers, typically 1-3 nm thick with heights of 0.2-0.5 eV, promoting resonant tunneling and . First demonstrated in 1994 using GaAs/AlGaAs heterostructures, QCLs achieve output powers over 1 W and operate up to , revolutionizing and sensing. As of 2025, advancements have enabled multi-watt outputs in compact devices for applications in free-space communications and high-resolution . In , the tip-sample interaction is approximated by a step potential barrier across the gap, typically 0.1-1 nm wide with a height of 4-5 eV, governing the tunneling current as a function of bias voltage. This model, refined in the Tersoff-Hamann approach, treats the tip as a and the sample's local as probing the evanescent wave decay, enabling atomic-resolution imaging and . Developed in the , STM has mapped surface electronic structures with sub-angstrom precision, influencing and . Advances since the early 2000s have explored step potential analogs in p-n junctions, where electrostatic gating creates sharp potential steps of 0.1-1 eV, manifesting Klein tunneling with near-perfect transmission at normal incidence due to the massless Dirac fermion description. Experimental evidence from transport measurements shows minimal resistance across such steps, contrasting non-relativistic backscattering and enabling pseudo-relativistic analogs in solid-state devices. This phenomenon, observed in gated structures, holds promise for high-speed transistors and .

References

  1. [1]
    [PDF] Scattering States and the Step Potential - MIT OpenCourseWare
    We now begin our detailed study of scattering states. These are un-normalizable energy eigenstates. They simply cannot be normalized, just like momentum ...
  2. [2]
    [PDF] 4.4 The Potential Step - Hansraj College
    So the motion of the particles is reversed by the potential barrier. Quantum mechanically, the picture will be somewhat different. In this case, the Schrödinger.
  3. [3]
  4. [4]
    [PDF] 14 Step potential and quantum tunneling - Physics Courses
    14.1 Step potential. The next system that we discuss quantum mechanically is the step potential, that is the potential energy of a particle is constant but ...Missing: definition | Show results with:definition
  5. [5]
    [PDF] Potential Step Consider a potential step described by1 V (x) = V0 for ...
    In quantum mechanics the particle has some probability (in an ensemble sense) to be reflected even though its energy is higher than the barrier. One way to ...
  6. [6]
    [PDF] Quantum mechanics applied to heterostructures
    Quantum mechanics for heterostructures. Two powerful, simplifying approximations. • Effective Mass Theory (the most important concept in semiconductor ...
  7. [7]
    [PDF] Physics 386
    The step potential​​ x. x=0. V0. There are several simple one-dimensional potentials that are of interest because they. approximate certain real physical ...
  8. [8]
    [PDF] Potential Step: Griffiths Problem 2.33 Prelude: Note that the time ...
    Potential Step: Griffiths Problem 2.33. Prelude: Note that the time-independent Schrödinger equation in one dimension can be written as d2ψ dx2. = −. 2m. ¯h2.
  9. [9]
    The Potential Step
    We will not require normalization to one particle, since we have a beam with definite momentum, which cannot be so normalized. (A more physical problem to ...Missing: unit | Show results with:unit
  10. [10]
    [PDF] Quantum Physics I, Lecture Note 6 - MIT OpenCourseWare
    Feb 23, 2016 · Ψ(x, t)+Ψ∗ Ψ(x, t) = (x, t) dx . ˆ Equation (2.11) is the condition on the Hamiltonian operator H for conservation of probability. Here we have ...
  11. [11]
    [PDF] Reflection and Transmission - Potential Step - MIT OpenCourseWare
    Question: What will T be if we double the width of the gap? • Let's consider a tunneling problem: An electron with a total energy of E o.
  12. [12]
    [PDF] Quantum Mechanics - IFSC/USP
    2.3.3 Quantum reflection at a potential step. The quantum reflection is a non-classical property of the motion of a particle. An example is the reflection of ...<|separator|>
  13. [13]
    [PDF] WENTZEL–KRAMERS–BRILLOUIN (WKB) APPROXIMATION
    • The bottom line is, the WKB approximation (27) is valid for a continuous potential ... discussion of the step potential in class and of the square ...
  14. [14]
    [PDF] Chapter 17: Resonant transmission and Ramsauer–Townsend
    Apr 26, 2016 · Ramsauer and Townsend reported a very unusual phenomenon. At very low energies, the scattering cross section was high.Missing: step | Show results with:step
  15. [15]
    [PDF] Chapter 4 Time–Independent Schrödinger Equation
    Definition 4.2 The parity operator P acting on a function f(x) changes the sign of its argument: P f(x) = f(−x).Missing: paper | Show results with:paper
  16. [16]
    [PDF] region region II
    The reflection and transmission coefficients happen to be given by the same expressions as what we had for the positive potential step. R = 0 .
  17. [17]
    [PDF] chapter 4
    For a particle that encounters a vertical step potential of height V0 , calculate the reflection and transmission coefficients for the particle energies. (a) E ...
  18. [18]
    Die Reflexion von Elektronen an einem Potentialsprung nach der ...
    Es wird die Reflexion von Elektronen an einem Potentialsprung nach der neuen Diracschen Dynamik untersucht. Bei sehr großen Werten des Potentialsprung.Missing: Oskar | Show results with:Oskar
  19. [19]
    [PDF] On the Klein's paradox in the presence of a scalar potential - arXiv
    Aug 1, 2019 · It is shown that the presence of a scalar barrier the scalar potential wides the gap between positive and negative energies and so the forbidden.
  20. [20]
  21. [21]
  22. [22]
    Resonant states and transmission coefficient oscillations for ...
    Apr 1, 2010 · Transmission across a rectangular barrier is the most commonly studied example in introductory quantum mechanics. In Fig. 3 we show the ...
  23. [23]
    A potential model for alpha decay | American Journal of Physics
    Sep 1, 2010 · We calculate the half-life and energy of alpha decay using a simple potential model consisting of the sum of the electrostatic Coulomb ...
  24. [24]
    Quantum phenomena explored with neutrons - IOPscience
    Neutrons are suitable tools for quantum experiments since they are massive, experience nuclear, electromagnetic and gravitational interaction and are easy to ...<|control11|><|separator|>
  25. [25]
    Computational issues in the simulation of semiconductor quantum ...
    2 the potential energy step at GaAs–AlGaAs or Si–SiO2 heterojunctions ... heterojunction step potential, n the quantum electron density, and finally V ...
  26. [26]
    Anisotropic transport of two-dimensional electron gas modulated by ...
    Jan 19, 2011 · ... heterojunction, with a tilted GaAs substrate13 and an in-plane ... step potential and anisotropic effective mass distribution. Although ...
  27. [27]
    [PDF] arXiv:cond-mat/0302341v1 [cond-mat.mes-hall] 17 Feb 2003
    radius a, with a very high step-potential at the edge r = a. ... For an AlGaAs/GaAs heterojunction, we choose the z-direction pointing from the AlGaAs to.
  28. [28]
    Intersubband electron transition across a staircase potential ...
    We present a theoretical investigation of a novel staircase-like light emitter based on the GaAs/GaxAl1−xAs material system. The emission wavelength is ...
  29. [29]
    Resonant tunneling times in superlattice structures
    arbitrary step potential we evaluate the coefficients of right- and left ... The quantum cascade laser, reported by Faist et al. ~Ref. 1!. ~a! The.
  30. [30]
    Open-boundary reflection of quantum well states at Pb(111)
    Nov 20, 2017 · Scanning tunneling microscopy. Article Text. I. INTRODUCTION ... step potential. Here we supplement that model, which permitted an ...
  31. [31]
    A simple method for simulating scanning tunneling images
    Aug 1, 2005 · In particular, we show how to model the apex of a STM as a source (or sink) of electrons, calculate three-dimensional tunneling across a vacuum ...
  32. [32]
    Evidence for Klein Tunneling in Graphene Junctions | Phys. Rev. Lett.
    Jan 16, 2009 · The Klein paradox for massless Dirac fermions predicts that carriers in graphene hitting a potential step at normal incidence transmit with ...Missing: relativistic analog 2000
  33. [33]
    Evidence of Klein tunneling in graphene p-n junctions - arXiv
    Jun 14, 2008 · Transport through potential barriers in graphene is investigated using a set of metallic gates capacitively coupled to graphene to modulate the potential ...Missing: relativistic analog post- 2000
  34. [34]
    (PDF) Evidence for Klein Tunneling in Graphene p-n Junctions
    Aug 6, 2025 · When a gate-induced potential step is steep enough, disorder becomes less important and the resistance across the step is in quantitative ...Missing: post- 2000