Fact-checked by Grok 2 weeks ago

Transmission coefficient

The transmission coefficient, denoted as T, is a in physics that represents the ratio of the transmitted , , or probability of a wave or particle to the corresponding incident value upon encountering a , , or potential barrier. This parameter is central to understanding wave propagation and in classical systems, where it arises from conditions at boundaries between media with differing properties, such as or impedance, ensuring such that T + R = 1, with R being the . In and acoustics, for instance, T determines the efficiency of or transmission through lenses, fibers, or layered materials, often expressed as T = \frac{2 Z_2}{Z_1 + Z_2} for amplitude transmission between two media with impedances Z_1 and Z_2. In quantum mechanics, the transmission coefficient gains particular significance for particles like electrons, where it quantifies the tunneling probability through classically forbidden regions, such as potential steps or barriers, even when the particle's energy E is below the barrier height V_0. For a step potential with E > V_0, it is given by T = \frac{4 k \bar{k}}{(k + \bar{k})^2}, where k = \sqrt{2mE}/\hbar and \bar{k} = \sqrt{2m(E - V_0)}/\hbar, with m the particle mass and \hbar the reduced Planck's constant; while for finite potential barriers with E < V_0, T is classically zero but finite quantum mechanically due to tunneling and wavefunction evanescence in the forbidden region. Beyond these core areas, the transmission coefficient informs diverse applications, including semiconductor tunneling in transistors, alpha decay in nuclear physics, and electromagnetic wave behavior in telecommunications antennas.

General Concepts

Definition

The transmission coefficient, often denoted as T, is defined as the ratio of the flux of the transmitted wave or particle to the incident flux upon interaction with a boundary or barrier. This flux typically represents energy, intensity, or probability current. Separately, there is an amplitude transmission coefficient, which is the ratio of the transmitted wave's amplitude to the incident wave's amplitude. These coefficients are applicable to both classical waves—such as light or sound—and quantum mechanical particles. The concept originated in early 19th-century wave optics, where Augustin-Jean Fresnel derived the transmission coefficients for light at dielectric interfaces in his 1823 memoir on light polarization and reflection. It was later generalized to quantum contexts in the 20th century, with George Gamow employing it in 1928 to describe the tunneling probability of alpha particles through nuclear potential barriers, explaining radioactive decay rates. As a pure ratio, the transmission coefficient is inherently dimensionless and, in passive systems lacking amplification, ranges from 0 (complete reflection) to 1 (complete transmission), quantifying the efficiency of passage across the boundary. The coefficient arises specifically under boundary conditions at interfaces between media with differing properties—such as acoustic impedance or refractive index—or at potential barriers, where continuity of the wave function and its derivative ensures partial reflection alongside transmission.

Properties and Interpretations

The transmission coefficient can be interpreted in different ways depending on the context. In classical wave phenomena, it often refers to the amplitude transmission coefficient, which is the ratio of the transmitted wave's field amplitude to the incident wave's amplitude. Alternatively, it describes the intensity or power transmission coefficient, representing the ratio of transmitted energy flux to incident energy flux. In quantum mechanics, the transmission coefficient corresponds to the probability that a particle passes through a potential region, given by the square of the modulus of the transmission amplitude. A fundamental property of the transmission coefficient in lossless systems is the conservation of energy, expressed as T + R = 1, where T is the transmission coefficient and R is the reflection coefficient. In time-reversal symmetric systems, reciprocity holds, such that the transmission coefficient from medium 1 to 2 equals that from 2 to 1, T_{12} = T_{21}. In quantum scattering theory, unitarity of the S-matrix ensures conservation of probability, leading to T + R = 1 for one-dimensional potentials. The value of the transmission coefficient depends on several factors, including impedance mismatch between media, which increases reflection and reduces transmission. It is also influenced by barrier height and width in potential scattering scenarios, wavelength or frequency relative to structural scales, and angle of incidence for oblique propagation. In classical wave mechanics, transmission vanishes below a cutoff frequency, such as in waveguides where waves become evanescent and do not propagate. In contrast, quantum mechanics permits tunneling, allowing nonzero transmission probabilities even when particle energy is below the barrier height, defying classical expectations.

Classical Wave Applications

Optics

In optics, the transmission coefficient refers to the transmittance T, defined as the ratio of transmitted light intensity I_t to incident light intensity I_i, i.e., T = I_t / I_i, representing the fraction of electromagnetic wave energy passing through an interface or medium without absorption or scattering. This quantity is crucial for understanding light propagation across boundaries between dielectrics, such as air-glass surfaces or layered materials, where it quantifies energy conservation in classical wave theory. The transmittance at an interface is governed by the Fresnel equations, which describe reflection and transmission for plane waves. For normal incidence on a non-absorbing interface, the amplitude transmission coefficient t is given by t = \frac{2 n_1}{n_1 + n_2}, where n_1 and n_2 are the refractive indices of the incident and transmitting media, respectively. The corresponding intensity transmittance T accounts for the change in wave impedance and is T = \frac{n_2}{n_1} |t|^2 = \frac{4 n_1 n_2}{(n_1 + n_2)^2}. This formula shows that T approaches 1 for matched indices (n_1 \approx n_2) but drops for mismatches, such as air (n_1 = 1) to glass (n_2 = 1.5), yielding T \approx 0.96 due to 4% reflection loss per surface. In absorbing or scattering media, T is further reduced by exponential decay terms, emphasizing the role of material properties in optical design. Applications of the transmission coefficient span optical devices relying on controlled light passage. In thin films, multilayer stacks engineer T via constructive or destructive interference; for instance, anti-reflection coatings on lenses use quarter-wave layers of intermediate-index materials to suppress reflection, boosting T from ~96% (uncoated glass) to over 99.5% at design wavelengths like 1064 nm. Optical filters exploit wavelength-dependent T: a blue bandpass filter achieves near-unity T (~90-95%) for blue light (~450 nm) while T \approx 0 (optical density >4) for red wavelengths (~650 nm), enabling color separation in imaging systems. in dyes or from particulates can diminish T across spectra, necessitating precise material selection for high-performance like camera lenses or solar cells. Transmittance is experimentally determined via , where a double-beam instrument measures I_t and I_i across UV-Vis-NIR wavelengths (e.g., 250-2500 nm) with accuracies of ±0.3% in the visible range. Samples like polished blocks are scanned to map T(\lambda), revealing effects from interfaces or internal losses; for a glass-air at normal incidence, spectrophotometric data confirms T \approx 0.96, aligning with Fresnel predictions after correcting for surface reflections.

Acoustics and Electromagnetic Waves

In acoustics, the transmission coefficient quantifies the fraction of incident that passes through an between two media, crucial for understanding wave propagation in engineering applications such as and . For normal incidence between two fluids, the intensity transmission coefficient is given by T = \frac{4 Z_1 Z_2}{(Z_1 + Z_2)^2}, where Z = \rho c denotes the characteristic , with \rho as the medium's density and c as the . This expression derives from the boundary conditions ensuring continuity of acoustic pressure and normal across the . At fluid-solid or air- interfaces, significant impedance mismatches often result in low transmission; for instance, air-water interfaces exhibit Z_\text{air} \approx 400 kg/m²s and Z_\text{water} \approx 1.5 \times 10^6 kg/m²s, yielding T \ll 1 and high , which is exploited in materials to attenuate noise by minimizing transmitted energy. A prominent example occurs in imaging, where the air-tissue poses a major challenge due to Z_\text{tissue} \approx 1.6 \times 10^6 kg/m²s, resulting in T \approx 0.001; this implies nearly 99.9% reflection, severely limiting signal penetration without coupling agents like gels that better match impedances and boost to practical levels. Such mismatches highlight the role of transmission coefficients in designing acoustic barriers for , where layered structures with graded impedances enhance overall by controlling reflections. For electromagnetic waves in non-optical regimes, such as radio frequencies and microwaves, transmission coefficients are essential in transmission lines for , where they describe the efficiency of or voltage transfer from source to load. The voltage transmission coefficient across an interface or load is T = \frac{2 Z_L}{Z_S + Z_L}, with Z_S as the source impedance and Z_L as the load impedance, while the reflection coefficient is \Gamma = \frac{Z_L - Z_S}{Z_L + Z_S}; optimal matching (Z_S = Z_L) yields T = 1 and \Gamma = 0, maximizing delivered . In , mismatches lead to return losses; for coaxial cables in cable TV systems with typical Z_0 = 75 \, \Omega, connecting to a 50 \Omega device produces \Gamma \approx 0.2, resulting in about 4% reflection and reduced . These principles underpin applications in antennas, where matching networks ensure efficient radiation, and in circuits to prevent echoes and distortion. At oblique incidence on dielectric interfaces, polarization effects influence transmission; for p-polarized waves (electric field parallel to the incidence plane), the reflection coefficient is zero at the Brewster angle \theta_B = \tan^{-1} \sqrt{\epsilon_2 / \epsilon_1}, where \epsilon are the permittivities, achieving T = 1 for power transmission without loss. This phenomenon, applicable in microwave and RF engineering, aids in designing polarization-selective devices like filters or absorbers. In contemporary 5G networks operating at millimeter waves, optimizing transmission coefficients via advanced impedance matching—such as adaptive circuits and metamaterial tuners—minimizes signal attenuation in base stations and user equipment to support higher data rates and coverage.

Quantum Mechanical Applications

Formulation

In quantum mechanics, the transmission coefficient arises in the study of one-dimensional scattering of particles by a potential, where the wave function satisfies the time-independent Schrödinger equation -\frac{\hbar^2}{2m} \frac{d^2 \psi(x)}{dx^2} + V(x) \psi(x) = E \psi(x), with E > 0 for scattering states and V(x) a potential localized such that V(x) \to 0 as |x| \to \infty. Solutions are sought in the form of stationary states representing plane waves incident from one side, with the wave function asymptotically approaching \psi(x) \to e^{i k_{\rm in} x} + r e^{-i k_{\rm in} x} as x \to -\infty (incident plus reflected waves) and \psi(x) \to t e^{i k_{\rm out} x} as x \to +\infty (transmitted wave), where k_{\rm in} = \sqrt{2mE}/\hbar and k_{\rm out} = \sqrt{2m(E - V_0)}/\hbar if the potential asymptotes differ by a constant V_0. Here, r and t are complex amplitudes encoding reflection and transmission, respectively. The coefficient T quantifies the probability of transmission and is defined as the ratio of the transmitted to the incident , T = J_{\rm trans}/J_{\rm inc}. The density in one dimension is given by J(x) = \frac{\hbar}{2mi} \left( \psi^* \frac{d\psi}{dx} - \psi \frac{d\psi^*}{dx} \right), derived from the \partial \rho / \partial t + \partial J / \partial x = 0 ensuring probability conservation, where \rho = |\psi|^2 is the probability density. For the asymptotic plane , the incident current is J_{\rm inc} = \hbar k_{\rm in}/m (with unit incident amplitude), while the transmitted current is J_{\rm trans} = |t|^2 (\hbar k_{\rm out}/m), yielding T = |t|^2 (k_{\rm out}/k_{\rm in}) to account for differing group velocities on each side. This formulation builds on classical transmission, which uses ratios of amplitudes, but introduces quantum evanescent (exponentially decaying solutions) in classically forbidden regions where E < V(x), enabling tunneling effects absent in deterministic classical descriptions. Scattering solutions are often organized using the S-matrix, which relates outgoing wave amplitudes to incoming ones via \begin{pmatrix} t & r' \\ r & t' \end{pmatrix}, where primes denote incidence from the right, connecting to the asymptotic forms \psi(x) \to t' e^{-i k_{\rm out} x} as x \to -\infty and \psi(x) \to e^{-i k_{\rm in} x} + r' e^{i k_{\rm in} x} as x \to +\infty. Unitarity of the , S^\dagger S = I, follows from current conservation and ensures T + R = 1, where R = |r|^2 (k_{\rm in}/k_{\rm in}) is the reflection probability (adjusted similarly for velocity if needed), reflecting the unitarity of time evolution in . For potentials invariant under time reversal and parity, t = t' and relations like r' = -r simplify the matrix further, but the general form captures asymmetric scattering.

Tunneling Through Barriers

In quantum mechanics, the transmission coefficient T represents the probability for a particle incident on a potential barrier to tunnel through it, even when the particle's energy E is less than the barrier height V_0. Classically, transmission would be impossible in this regime, but the particle's wave function extends into the forbidden region, decaying exponentially as e^{-\kappa x} for x > 0, where \kappa = \sqrt{2m(V_0 - E)} / [\hbar](/page/H-bar), m is the particle , and \hbar is the reduced Planck's ; this evanescent yields a non-zero T determined by matching the wave function and its derivative at the boundaries. A foundational example is , where George Gamow's 1928 theory modeled the escape of an from a heavy as tunneling through the . The transmission coefficient takes the approximate form T \approx e^{-2\pi \eta}, with \eta = \frac{2\pi Z_\alpha Z_d e^2}{\hbar v} the Sommerfeld parameter, Z_\alpha = 2 and Z_d the atomic numbers of the and daughter , e the , and v the velocity near the nuclear surface; this exponential factor explains the observed decay rates across isotopes. For the paradigmatic rectangular barrier of height V_0 and width L, the exact transmission coefficient when E < V_0 is T = \left[ 1 + \frac{V_0^2 \sinh^2(\kappa L)}{4 E (V_0 - E)} \right]^{-1}, where \kappa = \sqrt{2m(V_0 - E)} / \hbar. In the low-energy, thick-barrier limit (\kappa L \gg 1), this simplifies to T \approx \frac{16 E (V_0 - E)}{V_0^2} e^{-2 \kappa L}, highlighting the dominant exponential decay with barrier width and height difference. Conversely, in the high-energy or thin-barrier limit (\kappa L \ll 1), \sinh(\kappa L) \approx \kappa L, so T \approx \left[ 1 + \frac{V_0^2 (\kappa L)^2}{4 E (V_0 - E)} \right]^{-1}, which approaches 1 as the barrier effectively vanishes. This tunneling mechanism underpins key technologies, such as the scanning tunneling microscope (STM) developed by Binnig and Rohrer in 1982, where electrons tunnel between a metallic tip and sample across a vacuum barrier of ~1 nm; the tunneling current I \propto T \propto e^{-2 \kappa d} (with d the tip-sample separation) provides atomic-scale surface topography by maintaining constant current via piezoelectric adjustments. Field emission relies on similar principles, as described in the 1928 Fowler-Nordheim theory, where a strong electric field F tilts the surface potential barrier into a triangular shape, enabling electron tunneling from a metal cathode; the emission current density follows J \propto F^2 \exp(-b / F), with the exponential incorporating the barrier transmission probability integrated over occupied states up to the Fermi level. In semiconductor devices, the tunnel diode invented by Esaki in 1958 exploits interband tunneling in degenerate p-n junctions, where heavy doping narrows the depletion region to ~10 nm, allowing valence electrons to tunnel directly to the conduction band; the forward current peaks when band alignment maximizes T, then declines due to misalignment, yielding negative differential resistance for high-speed switching. For multilayer barriers, such as in quantum wells or superlattices, the transfer matrix method efficiently computes T by constructing a product of 2×2 matrices, each relating the wave function coefficients on either side of a layer via boundary continuity, enabling precise modeling of resonant peaks where T \approx 1 at specific energies.

WKB Approximation

The Wentzel–Kramers–Brillouin (WKB) approximation is a semi-classical technique for obtaining approximate solutions to the one-dimensional time-independent Schrödinger equation, -\frac{\hbar^2}{2m} \frac{d^2 \psi}{dx^2} + V(x) \psi = E \psi, where the potential V(x) varies slowly with position x. This method is valid for scenarios where the local de Broglie wavelength \lambda(x) = h / \sqrt{2m |E - V(x)|} changes gradually, specifically satisfying the condition |d\lambda / dx| \ll 1. Equivalently, the local wave number k(x) = \sqrt{2m |E - V(x)|} / \hbar must have a small relative derivative, |dk/dx| / k^2 \ll 1. Under these conditions, the wave function in classically allowed regions (E > V(x)) takes the form \psi(x) \approx k(x)^{-1/2} \exp\left(\pm i \int^x k(x') dx'\right), while in forbidden regions (E < V(x)), it is \psi(x) \approx \kappa(x)^{-1/2} \exp\left(\pm \int^x \kappa(x') dx'\right), with \kappa(x) = \sqrt{2m (V(x) - E)} / \hbar. These asymptotic forms, derived from expanding the phase in powers of \hbar, enable matching across turning points via connection formulas involving Airy functions. For calculating the transmission coefficient through a smooth potential barrier, the WKB method connects the oscillatory solutions outside the barrier to the exponentially decaying solution inside. Let x_1 and x_2 (x_1 < x_2) be the classical turning points where V(x_1) = V(x_2) = E. The approximate transmission probability T, representing the probability of tunneling from left to right, is given by T \approx \left[1 + \exp(2 \gamma)\right]^{-1}, where \gamma = \int_{x_1}^{x_2} \frac{\sqrt{2m [V(x) - E]}}{\hbar} \, dx. This formula arises from applying the WKB connection formulas at both turning points, accounting for the phase shifts and amplitude matching; providing better accuracy than the leading-order exponential T \approx \exp(-2\gamma) for cases where \gamma is not extremely large. For highly opaque barriers (\gamma \gg 1), it simplifies to T \approx \exp(-2\gamma), highlighting the semi-classical tunneling exponent. The WKB approximation finds key applications in estimating alpha decay rates, where it models the transmission of alpha particles through the surrounding the nucleus. George Gamow applied this method to derive the exponential dependence of decay lifetimes on energy, successfully reproducing the empirical for radioactive decay constants across isotopes. In molecular physics, WKB is employed to compute tunneling contributions to vibrational energy levels in anharmonic potentials, such as those in diatomic molecules like H_2^+, where it predicts splitting in symmetric double-well potentials and aids in interpreting spectroscopic data for overtone and combination bands. These applications underscore its utility in systems with smooth, monotonically decreasing barriers. Despite its successes, the WKB approximation has limitations when the potential varies rapidly near turning points, such as in thin barriers, where the slow-variation condition |d\lambda/dx| \ll 1 is violated, leading to significant errors in the transmission coefficient—often overestimating or underestimating T by orders of magnitude compared to exact solutions. It also inadequately captures oscillations near the barrier top, where exact quantum transmission approaches 1/2 rather than the WKB prediction of unity. Improvements to the standard WKB have focused on higher-order corrections and extensions for broader energy ranges. For instance, Kemble's formulation refines the prefactor via analytic continuation into the complex plane, enhancing accuracy for sub- and near-barrier energies. Post-2014 numerical validations, including integrations with exact scattering matrix methods and wavefunction benchmarks for proton tunneling in molecular systems, confirm that these updated WKB variants achieve errors below 5% for smooth potentials in heavy-ion fusion and vibrational tunneling simulations, maintaining relevance in contemporary quantum dynamics studies as of 2025.

Applications in Chemical Kinetics

Transition State Theory

Transition state theory (TST), also known as absolute reaction rate theory, provides a framework for understanding chemical reaction rates by positing the existence of a high-energy transition state through which reactants must pass to form products. Developed independently in 1935 by and by and , this theory revolutionized chemical kinetics by offering a statistical mechanical basis for rate predictions. In TST, the reaction rate constant k is expressed as k = \frac{k_B T}{h} e^{-\Delta G^\ddagger / RT} \kappa, where k_B is Boltzmann's constant, T is temperature, h is Planck's constant, \Delta G^\ddagger is the Gibbs free energy of activation, R is the gas constant, and \kappa is the transmission coefficient that corrects for deviations from the ideal assumption of unidirectional passage through the transition state. The transmission coefficient \kappa typically accounts for classical recrossing trajectories, where reactive species cross the dividing surface but return to reactants (\kappa < 1), or for quantum tunneling effects that enhance barrier penetration (\kappa > 1); for many classical monomolecular reactions, \kappa \approx 1. Historically, the introduction of \kappa in Eyring's formulation addressed the overestimation of rates in the basic TST model by incorporating probabilistic factors for non-ideal barrier crossing, evolving from earlier calculations by Eyring and Polanyi in 1931. In bimolecular substitution reactions like SN2 processes, \kappa often deviates significantly from due to corner-cutting paths that allow trajectories to bypass the canonical . The transmission coefficient in TST bridges classical and quantum descriptions by extending the concept of one-dimensional barrier tunneling probabilities to multidimensional spaces, where quantum effects can enhance rates through tunneling despite the theory's primarily classical foundations.

Transmission Factor in Rate Constants

In , the transmission coefficient κ modifies the to account for dynamical effects beyond the ideal assumption of a one-way flux through the , such as trajectory recrossing of the dividing surface and quantum tunneling. The full for the rate constant k of a is given by k = \frac{k_B T}{h} \left( \frac{RT}{p^0} \right)^{\Delta n} \exp\left( -\frac{\Delta G^\ddagger}{RT} \right) \kappa, where k_B is the Boltzmann constant, h is Planck's constant, T is the temperature, R is the gas constant, p^0 is the standard pressure (typically 1 bar), \Delta n is the change in the number of moles (e.g., \Delta n = -1 for a bimolecular reaction), and \Delta G^\ddagger is the standard free energy of activation. In the simplest case with no recrossing or tunneling, κ equals 1, representing perfect transmission through the barrier; however, κ is often computed using advanced methods like variational transition state theory (VTST), which optimizes the dividing surface to minimize recrossing, or ring-polymer molecular dynamics (RPMD), a path-integral approach that incorporates quantum effects in multidimensional systems. Variations in κ arise primarily from quantum tunneling and multidimensional recrossing, particularly in reactions involving light atoms. For tunneling corrections, the Bell model provides a semiclassical estimate for parabolic barriers, yielding a tunneling transmission factor κ_tun ≈ exp(ΔE_tun / k_B T), where ΔE_tun represents the effective energy gain from tunneling below the classical barrier; this model is especially relevant for proton or transfers at low temperatures. In multidimensional cases, κ is evaluated from classical or quasiclassical trajectory simulations, capturing recrossing due to coupling between the and orthogonal modes; for enzyme-catalyzed reactions, such computations yield κ values ranging from less than 1 due to recrossing to tens or more due to tunneling, reflecting significant deviations from unity due to and protein dynamics. Hydrogen transfer reactions exemplify cases where κ exceeds 1, driven by tunneling that enhances rates beyond classical predictions; for instance, in the isomerization of malonaldehyde, tunneling contributions make κ > 1 at room temperature, accelerating the proton relay by factors of up to several orders of magnitude. Isotope effects further highlight tunneling's role, as deuterium substitution reduces κ compared to protium due to the heavier mass suppressing quantum delocalization; kinetic isotope effects (KIEs) on κ can reach values of 2–7 for H/D in hydrogen abstractions, aiding identification of tunneling mechanisms in experimental rate data. Recent computational advances since 2014 have improved κ accuracy through simulations, addressing limitations in earlier implementations by integrating RPMD with machine-learned potentials for efficient sampling of complex surfaces. For example, RPMD-based methods now compute κ for polyatomic reactions with errors below 10% relative to exact quantum benchmarks, even in the deep tunneling regime, enabling reliable predictions for astrochemical and biomolecular systems. theory, another post-2014 development, optimizes tunneling paths in high dimensions to refine κ_tun, particularly for barrierless or low-barrier transfers.

References

  1. [1]
    [PDF] chapter 4
    The classical transmission coefficient, denoted T , is the ratio of intensity transmitted to intensity incident. Quantum mechanically, intensity is analogous ...
  2. [2]
    9.1: Reflection and Transmission - Physics LibreTexts
    Aug 2, 2021 · Notice that we have defined the reflection and transmission coefficients by taking out a factor of the amplitude, \(A\), of the incoming wave.
  3. [3]
    [PDF] Scattering States and the Step Potential - MIT OpenCourseWare
    We also define the transmission coefficient T to be the ratio of the probability flux in the transmitted wave to the probability flux in the incoming wave ...
  4. [4]
  5. [5]
    Reflection and Transmission Coefficients - Waves - NDE-Ed.org
    Coefficients: A calculated measure of a physical property such as reflected or transmitted waves. R = ( Z 2 − ...
  6. [6]
    The Fresnel Laws - Wiley Online Library
    Con- fusion has arisen from time to time in the way that the Fresnel coefficients have been presented. Fresnel's original paper of 1823 (see p. 777 in Fresnel, ...
  7. [7]
    The Early History of Quantum Tunneling - Physics Today
    Aug 1, 2002 · Gurney and Condon evaluated the integral graphically, but Gamow—although averse to complex mathematical analysis—produced an excellent analytic ...
  8. [8]
    [PDF] Relation between Transmission and Energy Stored in Random ...
    Sep 1, 2010 · This is essentially an energy conservation condition. ... The reciprocity of the system makes the transmission coefficient the same in both cases.
  9. [9]
    Reciprocity-induced symmetry in the round-trip transmission through ...
    Reciprocity is a fundamental principle of wave physics and directly relates to the symmetry in the transmission through a system when interchanging the input ...
  10. [10]
    [PDF] 10. Scattering Theory - DAMTP
    Indeed, in quantum mechanics the word. “unitarity” is often used synonymously with “conservation of probability”. One further property follows from the fact ...
  11. [11]
    [PDF] (6.3) Transmission Through a Layer: Normal Incidence Applicable to
    ... Transmission Coefficients for oblique incidence: I. R. R π = but in ... coefficient for various ratios of propagation speed and characteristic impedance.
  12. [12]
    [PDF] Lecture-Rectangular-waveguide.pdf - EMPossible
    •Not a transmission line because it has only one conductor. •Does not support a TEM mode. •Exhibits a low-frequency cutoff below which no waves will propagate.
  13. [13]
    7.7: Quantum Tunneling of Particles through Potential Barriers
    Mar 16, 2025 · Quantum tunneling is a phenomenon in which particles penetrate a potential energy barrier with a height greater than the total energy of the particles.
  14. [14]
    Fresnel Equations - RP Photonics
    Fresnel equations are equations for the amplitude coefficients of transmission and reflection at the interface between two transparent homogeneous media.Missing: 1823 original
  15. [15]
    Fresnel Equations - The University of Arizona
    Developed in the years 1821-1823, the Fresnel equations[1] describe the amplitude of transmitted and reflected light at the boundary between two materials.Missing: source | Show results with:source
  16. [16]
    [PDF] TIE-35: Transmittance of optical glass
    The transmittance of optical glass is measured using double beam spectral photometers from the company Perkin Elmer with special modifications. The standard ...
  17. [17]
    An Introduction to Optical Coatings
    ### Summary of Applications of Transmission Coefficients in Thin Films, Anti-Reflection Coatings, and Optical Filters
  18. [18]
    Introduction to Bandpass Optical Filters - Idex-hs.com
    30-day returnsTransmission: The percentage of light, T, that passes through the filter passband. · Blocking: Measured as Optical Density (OD), defined as OD = – log10 (T).
  19. [19]
    An Introduction to Optical Coatings
    Sep 3, 2024 · Optical coatings are thin films applied to optical devices to enhance their performance by manipulating light interactions such as reflections, transmission, ...Optical Coating Theory · Types Of Optical Coatings · References And Further...<|control11|><|separator|>
  20. [20]
    [PDF] 1 Reflection and tranmission of sound at an inter- face - MIT
    Consider two semi-infinite fluids separated by the plane interface along z = 0. The lower fluid is distinguished from the upper fluid by the subscript ”1”.
  21. [21]
    Waves at interfaces — GPG 0.0.1 documentation - GeoSci.xyz
    We define a new quantity called acoustic impedance as Z = ρ V , the product of density and velocity. ... Transmission Coefficient:.
  22. [22]
    Intensity Reflection Coefficient - CIE A Level Physics - Save My Exams
    Dec 25, 2024 · Air, Z1 = 4.3 × 102 kg m-2 s · Soft tissue, Z2 = 1.6 × 106 kg m-2 s ...
  23. [23]
    [PDF] Transmission Lines 2 - UCSB ECE
    1. Transmission line with impedance Zo connects the source with Thevenin equivalent generator impedance Zs to a load ZL. Vs =Re Voe.Missing: Z_L Z_S
  24. [24]
    Brewster's angle - RP Photonics
    The Brewsters angle is an angle of incidence at which there is no reflection of p-polarized light at an uncoated optical surface.
  25. [25]
    Impedance Matching Optimization of RF Networks - MDPI
    Aug 12, 2024 · The subject of this publication is a synthesized description of an S-parameter linear analysis for the improvement of a particular RF amplifier.
  26. [26]
    [PDF] Quantum Mechanics in One Dimension, Part II - Texas Tech University
    Sep 17, 2020 · Quantum Mechanics ... These coefficients are defined as the ratios of the probability fluxes, e.g., the transmission coefficient T is jII/jI,in.<|control11|><|separator|>
  27. [27]
    [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 ...
  28. [28]
    Schrödinger Wave Equation - Richard Fitzpatrick
    ... equation for the probability density: $\displaystyle \frac{\partial \rho}{\partial t}, (284). The probability current, $ {\bf j}$ , takes the form ...
  29. [29]
    [PDF] Physics 386
    , and quantum mechanics tells us that there is transmission through the ... This transmission coefficient is a tunneling probability for a single electron.
  30. [30]
    [PDF] 3. Scattering, Tunneling and Alpha Decay - MIT OpenCourseWare
    The difference here is that a finite transmission through the barrier is possible and the transmission coefficient is not zero. ... where G is the so-called Gamow ...
  31. [31]
    Surface Studies by Scanning Tunneling Microscopy | Phys. Rev. Lett.
    Jul 5, 1982 · This collection marks the 35th anniversary of scanning tunneling microscopy (STM) and the 30th anniversary of atomic force microscopy (AFM).
  32. [32]
    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.
  33. [33]
    [PDF] Leo Esaki - Nobel Lecture
    7. Current-voltage characteristics in SnTe and GeTe tunnel junctions at 4.2 K. their derivatives at each boundary, then calculate the transmission coefficient.
  34. [34]
    Resonant tunneling in multilayer systems | Phys. Rev. B
    Nov 15, 1988 · In contrast to transfer-matrix methods, it has the advantage of being stable for all systems without requiring the truncation of the growing ...
  35. [35]
    [1612.03014] Improved WKB approximation for quantum tunneling
    Dec 9, 2016 · In this paper we revisit the one-dimensional tunneling problem. We consider Kemble's approximation for the transmission coefficient. We show how ...
  36. [36]
    Tunneling dissociation of H 2 + and its isotopes in THz laser pulses
    Sep 23, 2014 · Therefore before doing quantum simulations, we use the WKB approximation to estimate the time that a nuclear wave packet takes to tunnel through ...
  37. [37]
    The Activated Complex in Chemical Reactions - AIP Publishing
    The probability of the activated state is calculated using ordinary statistical mechanics. This probability multiplied by the rate of decomposition gives the ...
  38. [38]
    Global dynamics and transition state theories: Comparative study of ...
    Jul 9, 2008 · 4 The transmission coefficient κ was calculated to be 0.55 ± 0.05 for the SN2 reaction Cl− + CH3Cl → ClCH3 + Cl− in water, which means a ...
  39. [39]
    Chemical Reaction Rate Coefficients from Ring Polymer Molecular ...
    Sep 14, 2016 · This Feature Article presents an overview of the current status of ring polymer molecular dynamics (RPMD) rate theory.
  40. [40]
    Unified ring polymer molecular dynamics rate calculations for ...
    Jul 8, 2025 · The ring polymer molecular dynamics (RPMD) rate theory is an efficient and accurate method for estimating rate coefficients of chemical reactions affected by ...
  41. [41]
    Coupling of hydrogenic tunneling to active-site motion in the ... - PNAS
    The standard deviations of the transmission coefficients are given in Table 1 and are small compared with the effect of reaction-path curvature, indicating ...
  42. [42]
    [PDF] Semiclassical Multidimensional Tunnelling Calculations
    The tests against exact quantum mechanics are direct tests of the accuracy of the calculated rate constant that are of special interest because the transmission ...<|control11|><|separator|>
  43. [43]
    Rate coefficients and kinetic isotope effects of the X + CH4 → CH3 + ...
    Mar 1, 2013 · κ is the plateau value of the transmission coefficient. kQTST and kRPMD are the centroid density QTST rate coefficients and the ...
  44. [44]
    Ring Polymer Molecular Dynamics Rates for Hydrogen ...
    Feb 6, 2025 · The conventional theoretical approach to calculating rate coefficients of an elementary surface reaction is based on transition-state theory ( ...