Fact-checked by Grok 2 weeks ago

Matter wave

A matter wave, also known as a de Broglie wave, refers to the wave-like properties associated with particles of matter, embodying the wave-particle duality fundamental to . In 1924, French physicist proposed that every particle or object with p possesses a \lambda = h / p, where h is Planck's constant, extending the wave nature previously observed in light to all matter. This hypothesis complemented Albert Einstein's 1905 explanation of the , which treated light as particles (photons), and suggested a symmetric duality where particles could interfere and diffract like waves. De Broglie's hypothesis paved the way for the development of wave mechanics by in 1926. The relation was experimentally verified in through the Davisson-Germer experiment, in which electrons scattered off a produced diffraction patterns consistent with a of approximately 1.65 for electrons accelerated at 54 eV, aligning closely with theoretical predictions. This confirmation earned de Broglie the in 1929. Matter waves have since been demonstrated for a wide range of particles, including protons, neutrons, and atomic nuclei, through similar and experiments. Further advancements have extended observations to neutral atoms and even complex molecules; for instance, in 2019, and were achieved with molecules containing up to 2000 atoms and a exceeding 25,000 atomic mass units, confirming the de Broglie wavelength scales inversely with and . These phenomena underpin applications in electron microscopy, atom interferometry for precision measurements, and technologies like , highlighting the profound impact of matter waves on and .

History

Background Concepts

At the turn of the 20th century, classical physics faced significant challenges in explaining certain experimental observations, particularly those involving radiation and energy emission. The blackbody radiation problem, known as the ultraviolet catastrophe, arose because classical Rayleigh-Jeans law predicted infinite energy at high frequencies, contradicting experimental spectra. In 1900, Max Planck resolved this anomaly by introducing the concept of energy quanta, proposing that oscillators in a blackbody emit and absorb energy in discrete units rather than continuously, which matched observed data and marked the birth of quantum theory. Building on Planck's ideas, extended the quantum hypothesis to itself in 1905, explaining the where ejects electrons from metals only above a threshold , regardless of intensity. Einstein posited that consists of discrete energy packets, or quanta (later called photons), with energy proportional to , thus attributing particle-like properties to electromagnetic . This challenged the purely wavelike view of from classical . Further evidence for light's particle nature came in 1923 with Arthur Holly Compton's experiments, where X-rays interacting with electrons produced wavelength shifts consistent with particle collisions, as if photons transferred momentum to electrons like billiard balls. These developments highlighted a growing wave-particle duality for light, where phenomena required both wavelike and particle-like descriptions depending on the context. Louis de Broglie, working in this intellectual climate, was influenced by Einstein's of and the quantum ideas of Planck and his brother Maurice de Broglie's research. In his 1924 doctoral thesis, de Broglie sought to extend duality to particles, proposing that if waves could behave as particles, particles might exhibit wave properties—a hypothesis that culminated these conceptual shifts.

De Broglie Hypothesis

In 1924, proposed that particles of matter, such as electrons, exhibit wave-like properties in addition to their particle nature, extending the wave-particle duality originally observed in light to all matter. This hypothesis posited that every moving particle is associated with a wave, which he termed a "pilot wave," guiding the particle's trajectory and providing a physical interpretation for quantum phenomena. Building on Einstein's light quanta and Planck's quantization, de Broglie argued that this duality resolves inconsistencies in early quantum models, such as Bohr's atomic theory, by attributing wave characteristics to electrons in stable orbits. Central to de Broglie's hypothesis is the relation between the wavelength \lambda of the associated matter wave and the particle's momentum p, given by \lambda = \frac{h}{p}, where h is Planck's constant. This relation derives from the photon case, where energy E = h \nu (with \nu the frequency) and momentum p = \frac{h}{\lambda} follow from the relativistic energy-momentum equivalence E = p c for massless particles; de Broglie generalized it to massive particles by assuming the same form holds, with the wave frequency linked to energy via E = h f. For a particle at rest, the frequency f corresponds to its rest energy m c^2, while motion introduces a phase wave with the derived wavelength. These relations were first outlined in de Broglie's doctoral thesis, supervised by Paul Langevin and defended on November 25, 1924, with preliminary ideas published in notes to the Comptes Rendus in 1922 and 1923. The hypothesis received mixed initial reception but gained early support from prominent physicists. , upon reviewing the thesis, expressed great enthusiasm in a letter to de Broglie's thesis supervisor on December 16, 1924, describing it as a profound contribution that uncovered "a corner of the great veil," which helped elevate the proposal's visibility among prominent physicists. This endorsement helped elevate the proposal's visibility, though broader acceptance awaited mathematical formalization, such as in the later . The full thesis was published in 1925.

Schrödinger Equation Development

In 1926, Erwin , building on Louis de Broglie's hypothesis that particles possess wave-like properties, formulated wave mechanics as a comprehensive framework for quantum phenomena. This development addressed limitations in the existing by providing a description of matter waves, transforming de Broglie's qualitative idea into a quantitative theory. 's approach was motivated during a holiday in the Austrian Alps, where he sought to extend classical analogies to quantum systems. Schrödinger published a series of four key papers in that year, beginning with "Quantisierung als Eigenwertproblem" on March 13, 1926. In these works, he introduced the time-dependent , which describes the temporal evolution of the wave function \psi(\mathbf{r}, t): i \hbar \frac{\partial \psi}{\partial t} = \hat{H} \psi Here, \hat{H} = -\frac{\hbar^2}{2m} \nabla^2 + V(\mathbf{r}) is the operator, incorporating kinetic and terms, \hbar is the reduced Planck's constant, m is the particle mass, and V(\mathbf{r}) is the potential. For time-independent problems and stationary states, the equation simplifies to the eigenvalue form: \hat{H} \psi = E \psi where E represents the energy of the state, treating quantization as an eigenvalue problem. The derivation of this equation stemmed from de Broglie's matter wave concept and analogies between classical wave equations and mechanics. Schrödinger drew on Hamilton's optico-mechanical analogy, linking ray optics to particle trajectories via the Hamilton-Jacobi equation, where the principal function S satisfies \frac{\partial S}{\partial t} + H(\mathbf{r}, \nabla S) = 0. He postulated a plane wave solution \psi = A e^{i(S/\hbar)} for de Broglie waves, substituting into a classical wave equation and expanding in powers of \hbar to bridge the geometric optics limit (corresponding to the Hamilton-Jacobi equation) with quantum corrections from higher-order terms. This yielded the Schrödinger equation as the fundamental wave equation for matter. Applying the equation to the in his fourth paper, Schrödinger demonstrated that it precisely reproduced the observed spectral lines, validating the wave mechanical approach against empirical data.

Early Experimental Confirmations

The first experimental confirmation of matter waves came from the Davisson-Germer experiment in 1927, where and Lester Germer at Bell Laboratories observed of electrons scattered by a surface. They directed a beam of electrons with energies around 54 eV onto a polycrystalline target and detected a strong peak in the scattered intensity at a 50-degree , which corresponded to constructive from the planes spaced at approximately 0.215 nm. Calculating the from the and spacing using yielded λ ≈ 0.165 nm, which precisely matched the de Broglie prediction λ = h/p for electrons of that . Independently in the same year, George Paget Thomson and his collaborators at Imperial College London demonstrated electron diffraction through transmission experiments using thin polycrystalline films. By passing a beam of 20-40 keV electrons through films of celluloid or gold foil about 100 nm thick, they observed ring-like diffraction patterns on a fluorescent screen, indicative of wave interference from the atomic lattice. The measured ring diameters aligned with de Broglie wavelengths calculated from the electron velocities, providing further evidence of the wave nature of electrons. For their groundbreaking discoveries, Davisson and Thomson shared the 1937 Nobel Prize in Physics. Extending these results to neutral atoms, and Immanuel Estermann conducted the first experiments with atomic beams in 1930. They scattered a beam of atoms from a clean (LiF) crystal surface and observed peaks at angles consistent with the de Broglie wavelength for helium's and , around 0.06 nm for room-temperature atoms. Similar patterns were seen for molecules, confirming wave properties for composite particles. The wave nature of neutrons was verified in 1936 by Hans von Halban and Paul Preiswerk at the Curie Laboratory in . Using neutrons from a radium-beryllium source moderated by , they diffracted the beam through a rock salt crystal and detected intensity maxima at expected Bragg angles, yielding a of about 0.19 nm that matched the de Broglie formula for thermal neutrons with velocities around 2000 m/s. This independent confirmation by Mitchell and Powers further solidified the universality of matter waves.

Fundamental Properties

Wavelength and Momentum Relation

The de Broglie relations define the fundamental wave properties of massive particles, associating a \lambda with the particle's p and a f with its total E, through the equations \lambda = \frac{h}{p}, \quad f = \frac{E}{h}, where h = 6.626 \times 10^{-34} \, \mathrm{J \cdot s} is Planck's constant. These formulas apply to any particle with rest , generalizing the relations \lambda = c/f and E = h f to by positing that particles exhibit wave-like behavior with phase determined by their mechanical properties. For non-relativistic particles, the p = m v, where m is the and v is the speed, so the inversely scales with both speed and . The dependence of wavelength on momentum implies that matter waves for faster or heavier particles are shorter, making wave interference effects more pronounced for lighter, slower particles where \lambda approaches atomic scales. For example, an electron with 100 eV kinetic energy has momentum p = \sqrt{2 m_e E} \approx 5.39 \times 10^{-24} \, \mathrm{kg \cdot m/s} (using m_e = 9.109 \times 10^{-31} \, \mathrm{kg} and E = 1.602 \times 10^{-17} \, \mathrm{J}), yielding \lambda \approx 0.123 \, \mathrm{nm} (1.23 Å), comparable to interatomic distances in crystals. A proton at the same energy, with mass m_p \approx 1.673 \times 10^{-27} \, \mathrm{kg} (about 1836 times that of the electron), has p \approx 2.30 \times 10^{-22} \, \mathrm{kg \cdot m/s} and \lambda \approx 2.88 \times 10^{-12} \, \mathrm{m} (0.0288 Å), over 40 times shorter due to the mass ratio. For atoms, such as sodium (m \approx 3.82 \times 10^{-26} \, \mathrm{kg}) at room temperature (300 K), the thermal de Broglie wavelength is \lambda_\mathrm{th} = h / \sqrt{2 \pi m k T} \approx 0.021 \, \mathrm{nm} (0.21 Å), still on the order of atomic radii but enabling interferometry in dilute gases. This wavelength sets the characteristic scale for quantum wave packets, where the spatial extent \Delta x of the packet relates to the inverse of the momentum spread \Delta p, aligning with the Heisenberg \Delta x \Delta p \geq \hbar / 2 (with \hbar = h / 2\pi) that limits simultaneous precision in position and momentum measurements. In practical units for and , wavelengths are often expressed in angstroms (1 Å = 0.1 ), highlighting their relevance to nanoscale phenomena.

Phase and Group Velocities

In matter waves, the phase velocity v_p and group velocity v_g characterize different aspects of wave propagation associated with particles. The phase velocity is defined as v_p = \frac{\omega}{k}, where \omega is the angular frequency and k is the wave number; using the de Broglie relations E = \hbar \omega and p = \hbar k, it simplifies to v_p = \frac{E}{p}. For non-relativistic particles, where the total energy E \approx mc^2 + \frac{p^2}{2m} and particle speed v \ll c, this yields v_p \approx \frac{c^2}{v}, exceeding the speed of light c. The , v_g = \frac{d\omega}{dk}, corresponds to the of a localized and equals the classical . For a , the derived from the is \omega = \frac{\hbar k^2}{2m}, so v_g = \frac{d\omega}{dk} = \frac{\hbar k}{m} = \frac{p}{m}. This matches the particle's momentum-based speed v = \frac{p}{m}. These velocities emerge from the superposition of plane waves forming a wave packet, where the envelope propagates at v_g while individual phases move at v_p. De Broglie and Schrödinger identified v_g as the particle's velocity, resolving how the wave guides the particle without contradicting classical limits. Consider an accelerated to 54 eV in early experiments: its speed is approximately $4.4 \times 10^6 m/s, so v_g \approx 4.4 \times 10^6 m/s, aligning with the , while v_p \approx 2.0 \times 10^{10} m/s, greatly exceeding c. However, this superluminal carries no or , preserving .

Relativistic Extensions

In the relativistic framework, the de Broglie relations for matter waves are extended to incorporate , where the \lambda remains \lambda = h / p with p as the relativistic p = \gamma m v, and the \nu satisfies h \nu = E with E = \gamma m c^2, \gamma = 1 / \sqrt{1 - v^2/c^2}. These relations unify the particle's p^\mu = (E/c, \mathbf{p}) with the wave k^\mu = (\omega/c, \mathbf{k}) through p^\mu = \hbar k^\mu, ensuring Lorentz invariance in the description of matter waves. The Klein-Gordon equation serves as the relativistic analog to the non-relativistic for scalar particles, derived by quantizing the relativistic energy-momentum relation E^2 = p^2 c^2 + m^2 c^4. It takes the form \left( \square + \frac{m^2 c^2}{\hbar^2} \right) \psi = 0, where \square = \partial^\mu \partial_\mu is the d'Alembertian operator, providing a for spin-0 particles but suffering from issues like densities. For spin-1/2 particles like electrons, the addresses the shortcomings of the Klein-Gordon approach by incorporating spin, yielding i \hbar \frac{\partial \psi}{\partial t} = c \boldsymbol{\alpha} \cdot \mathbf{p} \psi + \beta m c^2 \psi, where \psi is a four-component , \boldsymbol{\alpha} and \beta are $4 \times 4 matrices, and this equation reduces to the non-relativistic in the low-speed limit. A key consequence of the is its prediction of through the existence of negative-energy solutions, interpreted via Dirac's 1930 hole theory as absences in a filled negative-energy sea, corresponding to particles with opposite charge. This prediction was experimentally confirmed in , first by Carl Anderson's discovery of the in 1932, and subsequently by and Giuseppe Occhialini's observation of electron-positron from interactions in 1933.

Types of Matter Waves

Single-Particle Matter Waves

In single-particle matter waves, the quantum behavior of an isolated particle, such as an or , is described by a \psi(\mathbf{r}, t), which serves as a for finding the particle at position \mathbf{r} at time t. The square of the modulus, |\psi(\mathbf{r}, t)|^2, gives the probability density of the particle's location, embodying the probabilistic nature of . This interpretation, introduced by in 1926, transforms the wave function from a mere descriptive tool into a fundamental predictor of outcomes. The time evolution of \psi(\mathbf{r}, t) is governed by the . A key manifestation of single-particle matter is self-, where the wave function splits and recombines, producing interference patterns characteristic of . In theoretical descriptions of the double-slit setup, an or atom's passes through both slits simultaneously, interfering with itself to yield a modulated on a distant screen, even if particles arrive singly. This underscores the particle's delocalized wave nature, with the interference fringes arising solely from the superposition of amplitudes from each path. Quantum tunneling exemplifies the wave's ability to penetrate barriers forbidden by . For instance, in , the alpha particle's matter wave extends into and beyond the potential barrier surrounding the , enabling probabilistic escape despite insufficient energy for classical transmission. This phenomenon, theoretically explained by in 1928, accounts for observed decay rates. The Aharonov-Bohm effect further highlights the subtle influence on matter wave phases: in 1959, and showed that the phase of an electron's can shift due to the electromagnetic , even when the particle traverses regions free of electric and magnetic fields. This phase shift manifests as altered patterns, emphasizing the non-local gauge-dependent aspects of single-particle waves. For isolated particles, the —the spatial extent over which the matter wave maintains —is inversely proportional to the momentum uncertainty \Delta p, roughly l_c \approx h / \Delta p, where h is Planck's constant. , caused by interactions with the environment that introduce random fluctuations, limits observable to distances shorter than l_c, degrading the wave's quantum .

Collective Matter Waves

Collective matter waves emerge in coherent ensembles of particles where quantum extends macroscopically, manifesting as a single, unified rather than individual particle behaviors. In such systems, interactions among many particles lead to nonlinear effects and collective excitations that propagate as matter waves, distinct from the linear superposition typical of single particles. A prime example is Bose-Einstein condensation (BEC), first achieved experimentally in 1995 using dilute gases of rubidium-87 atoms cooled to near at by Eric Cornell and . In BEC, a large number of bosonic atoms occupy the lowest , forming a macroscopic matter wave described by a single wave function ψ(r,t) whose square modulus represents the particle density. This condensate behaves as a coherent , enabling phenomena like matter wave interference on macroscopic scales. The 2001 was awarded to Cornell, Wieman, and for this achievement and subsequent studies of BEC properties in ultracold gases. Applications include precision measurements and quantum simulation using these coherent ensembles. The dynamics of BEC are governed by the Gross-Pitaevskii equation (GPE), a derived in the mean-field approximation for weakly interacting bosons. Independently proposed by Eugene Gross in and Lev Pitaevskii in , the time-dependent GPE reads: i \hbar \frac{\partial \psi}{\partial t} = \left[ -\frac{\hbar^2}{2m} \nabla^2 + V(\mathbf{r},t) + g |\psi|^2 \right] \psi where ħ is the reduced Planck's constant, m is the particle mass, V is the external potential, g is the interaction strength proportional to the s-wave scattering length, and ψ is the wave function normalized such that ∫|ψ|² d³r = N, the total number of particles. This equation captures collective matter wave propagation, including solitons and vortex formation in the . Earlier insights into collective matter waves came from , where proposed in 1938 that arises from Bose-Einstein degeneracy, allowing the liquid to behave as a macroscopic quantum wave. In below the λ-transition of 2.17 K, atoms form a that supports frictionless flow and waves—temperature oscillations propagating as collective matter waves without viscosity. This theory linked to matter wave coherence in dense fluids. Fermionic systems exhibit analogous collective matter waves through in paired states, where fermions form bosonic pairs that condense similarly to bosons. The first realization of a fermionic superfluid occurred in 2003-2004 with ultracold atoms near a Feshbach , enabling tunable pairing and unitary superfluidity. These paired states support collective excitations like sound waves and vortices, extending matter wave phenomena to fermions.

Standing Matter Waves

Standing matter waves form through the of two counter-propagating matter waves of equal and , producing a pattern characterized by fixed nodes (points of zero ) and antinodes (points of maximum ). This superposition results in a time-independent , where the remains constant across the pattern, contrasting with propagating waves that carry over distance. In bound quantum systems, such as the infinite square well potential, standing matter waves lead to quantized energy levels due to the boundary conditions requiring the wave function to vanish at the walls. For a particle of mass m confined to a one-dimensional box of width L, the energy eigenvalues are given by E_n = \frac{n^2 \pi^2 \hbar^2}{2 m L^2}, where n = 1, 2, 3, \dots is a positive integer labeling the states, and the corresponding wave functions are \psi_n(x) \sim \sin\left( \frac{n \pi x}{L} \right). These solutions arise because only wavelengths that fit an integer number of half-waves within the well satisfy the condition. This concept extends to atomic orbitals, where electrons in atoms form patterns around the , determining the spatial probability distributions and discrete energy levels in multi-electron systems. Similarly, in semiconductor —nanoscale structures that confine in three dimensions— produce size-dependent discrete energy levels, enabling applications in like quantum dot lasers. The Franck-Hertz experiment of 1914 provided early empirical support for such quantization by demonstrating discrete energy losses in electron-mercury atom collisions, prefiguring the role of standing waves in bound states. In potential barriers, matter waves transition to evanescent forms, where the wave function decays exponentially in the classically forbidden region without oscillating, facilitating quantum tunneling while maintaining the standing wave character at the boundaries.

Experimental Evidence

Electron Diffraction and Interference

The Davisson-Germer experiment in 1927 provided the first direct experimental confirmation of , demonstrating that electrons scatter from a surface in a manner consistent with , producing intensity maxima at specific angles corresponding to the spacing. In this setup, a beam of slow electrons (around 54 eV) was directed at a polycrystalline target, and the angular distribution of scattered electrons revealed peaks that matched the predictions for from atomic planes separated by 0.215 nm, aligning with the de Broglie calculated for those electrons. Building on this, (LEED) emerged in the mid-20th century as a refined technique for probing surface structures, using electrons with energies typically between 20 and 200 eV to generate patterns from ordered surfaces. patterns, observed on fluorescent screens or via electron detectors, exhibit spots whose positions and intensities reveal the two-dimensional arrangement of surface atoms, with the method's sensitivity to the top few atomic layers stemming from the short of low-energy electrons. This development extended the original Davisson-Germer approach by enabling quantitative analysis of surface reconstructions and adsorbate layers through kinematic and dynamical scattering theories. A landmark demonstration of electron wave interference came in 1961 with Claus Jönsson's double-slit experiment, where accelerated to 25-40 keV passed through two 0.3 μm wide slits separated by 1-2 μm, producing clear fringes on a distant screen with a spacing matching the expected pattern for de Broglie wavelengths of approximately 0.007 nm. This setup confirmed single-particle wave behavior without relying on crystal lattices, as the fringe visibility persisted even at low electron fluxes, indicating self- of individual . Modern iterations, such as the 2013 controlled double-slit experiment using a , further isolated single-electron events by mechanically opening and closing nanofabricated slits (40 nm wide, 3 μm separation) at 80 keV, reconstructing patterns from individual detections that built up statistically over time. Field emission microscopy has also revealed electron wave patterns through interference, particularly in setups where electrons emanate from closely spaced nanotips, such as multiwalled carbon nanotubes, generating striped emission fringes on a screen due to the coherent superposition of waves from adjacent emitters. In one such observation at 60 K, interference fringes with periods of about 0.5 nm arose from de Broglie wavelengths around 0.005 nm for s at several keV, directly visualizing the phase coherence over micrometer scales in the emission pattern. The de Broglie relation, λ = h / p, yields an wavelength of approximately 0.004 nm at 100 keV, setting fundamental limits on the achievable in electron-based and setups like . Recent experiments in the 2020s have advanced toward delayed-choice paradigms; for instance, a 2019 implementation using a biprism and nanofabricated slits allowed control of visibility for single 80 keV electrons, effectively demonstrating path information erasure post-scattering by adjusting beam deflections after electron passage.

Neutron and Proton Experiments

One of the earliest demonstrations of matter waves for neutrons came from diffraction experiments conducted by Immanuel Estermann, Otto C. Simpson, and Otto Stern in 1936, who directed a beam of slow neutrons onto crystalline targets such as rock salt and observed interference patterns consistent with de Broglie wave scattering, thereby verifying the wave-particle duality for neutrons and laying the groundwork for neutron-based crystal structure determination. These results showed diffraction maxima at angles predicted by Bragg's law for neutron wavelengths around 1 Å, highlighting the utility of neutrons' neutral charge and magnetic moment in probing atomic arrangements without significant electrostatic interference. Neutron emerged in the through the work of and collaborators at the Institut Laue-Langevin, who developed crystal-based interferometers to split and recombine beams, enabling precise measurements of shifts due to gravitational fields. A seminal experiment by Colella, Overhauser, and Werner in 1975 utilized such an interferometer to observe a gravitationally induced shift in traversing paths at different heights in Earth's , with the difference Δφ = (2π m g A cos θ) / (h v) matching predictions to within 1%, where m is , g is , A is interferometer area, v is , h is Planck's , and θ is the tilt . This confirmed the influence of on quantum matter waves, with subsequent Zeilinger-led refinements in the late achieving higher visibility contrasts up to 90% by optimizing beam coherence. The 1975 Colella et al. experiment used thermal s in a crystal interferometer to observe a gravitationally induced shift, matching predictions to within 1%. In the , advances in matter wave experiments have explored BEC-like coherent states using ultra-cold s confined in gravitational traps, where quantum bounces produce macroscopic wavefunctions exhibiting collective patterns akin to Bose-Einstein condensates, as demonstrated in qBounce setups testing modified theories with precisions below 0.1 radians. These developments, including entanglement imaging with beams, enhance sensitivity for searches via interferometric anomalies.

Atomic and Molecular Demonstrations

Experiments demonstrating matter wave interference with neutral atoms and small molecules have provided key evidence for the wave nature of composite particles. In the 1930s, Immanuel Estermann and conducted pioneering experiments using beams of atoms and molecular (H₂) incident on the surfaces of (LiF) or (NaCl) crystals. These observations confirmed the de Broglie relation for neutral atoms and diatomic molecules, with patterns matching predictions based on the molecular de Broglie derived from beam velocities around 1000–2000 m/s. The first direct observation of atomic matter wave occurred in 1991, when Mark Kasevich and demonstrated a light-pulse using laser-cooled sodium atoms. In this setup, stimulated Raman transitions served as beam splitters and mirrors, creating a Mach-Zehnder-like interferometer where atomic wave packets were split, redirected, and recombined, producing fringes with visibilities up to 30%. For these atoms, with center-of-mass velocities on the of several cm/s after , the de Broglie wavelength was approximately 1 μm, enabling coherent over path separations of several millimeters. To achieve such coherence, techniques were essential, reducing atomic thermal velocities and increasing the de Broglie wavelength while minimizing decoherence from . The development of these methods, recognized by the 1997 awarded to , , and William D. Phillips, allowed atoms to be cooled to microkelvin temperatures, enhancing matter wave phase stability for interferometric demonstrations. Atom interferometry has since advanced with setups using rubidium-87 atoms in Mach-Zehnder configurations, where sequences of Raman pulses split and recombine wave functions to measure shifts from gravitational or inertial effects. These experiments, often employing cold clouds from magneto-optical traps, achieve contrast visibilities exceeding 90% and sensitivities to differences on the order of milliradians, as reviewed in comprehensive surveys of atom optics techniques. For molecules, diffraction experiments evolved in the 1990s with the use of material gratings and supersonic beam sources to produce coherent beams of small molecules like helium dimers or simple diatomics. These setups demonstrated single-slit and grating diffraction patterns, verifying wave interference for molecular matter waves with de Broglie wavelengths around 0.1–1 nm, and paved the way for more complex interferometers by controlling molecular velocities to below 100 m/s. In the , hybrid atom-molecule interferometers have emerged, utilizing Feshbach resonances to reversibly associate ultracold atoms into molecules within the interferometer arms, allowing direct comparison of single-particle and bound-state matter wave behaviors. These trapped systems, often with fermionic or bosonic atoms forming weakly to strongly interacting Feshbach molecules, exhibit interference contrasts up to 80% and enable studies of interaction effects on matter wave propagation.

Recent Advances with Large Molecules

Recent advances in matter wave experiments have pushed the boundaries of to increasingly large and complex molecules, demonstrating wave-like for objects approaching macroscopic scales. A seminal demonstration occurred in 1999 when patterns were observed for C60 fullerenes, molecules with a of approximately 720 u, using a interferometer; the de Broglie wavelength was on the order of picometers, confirming wave-particle duality for these soccer-ball-shaped carbon cages. These experiments marked the first observation of matter wave for objects composed of 60 atoms, highlighting the universality of quantum behavior despite the molecules' size and internal complexity. Building on this foundation, experiments in 2019 extended interference observations to functionalized oligoporphyrin molecules, including derivatives of and phthalocyanine-like structures, with masses exceeding 25,000 Da and up to 2,000 atoms. Using a Talbot-Lau interferometer, researchers achieved high-contrast fringe patterns with visibilities up to 90%, even at de Broglie wavelengths as small as 53 femtometers, scaled according to the de Broglie relation λ = h / p for these massive particles. These results pushed the quantum-classical boundary, serving as analogs to by placing the entire molecular in superposition over nanoscale separations. Further progress with biomolecules came in 2020, when matter wave interference was demonstrated for , a native consisting of 15 (mass ~1,881 , roughly 100 atoms total), achieving fringe visibilities exceeding 90% in an environment. This experiment approached decoherence limits by isolating the fragile from environmental interactions, confirming quantum delocalization over distances more than 20 times its molecular diameter. Such tests with biomolecules underscore the potential for quantum effects in biological systems, though challenges persist, including rapid decoherence due to thermal vibrations and collisions with residual gas molecules, which are mitigated through cryogenic cooling to millikelvin temperatures and levels below 10-10 mbar. Extending beyond molecules, matter-wave has been achieved with nanoparticles in the 2020s. In 2025, experiments demonstrated of nanoparticles with masses exceeding 106 units, using advanced beam sources and interferometers to probe the quantum-classical boundary at larger scales.

Comparisons with Other Waves

Differences from Electromagnetic Waves

Matter waves, unlike electromagnetic waves, are intrinsically linked to particles with non-zero rest mass, such as electrons or atoms, while electromagnetic waves consist of massless photons. This mass distinction profoundly affects their propagation: the of matter waves corresponds to the classical velocity of the associated particle, which is always subluminal (less than the c), whereas the of electromagnetic waves in is precisely c, independent of . A key consequence of this is the dispersive nature of matter waves, where the vp differs from the vg; in the non-relativistic case, vp = vg / 2, while in the relativistic case, vp = c2 / vg exceeding c, leading to spreading over time. In stark contrast, electromagnetic in are non-dispersive, maintaining vp = vg = c for all wavelengths, ensuring undistorted without spreading. waves thus obey the non-relativistic for description, while photons require the relativistic framework of (QED), precluding a direct Schrödinger-like equation for their wave behavior. Detection methods further highlight these differences: matter waves are observed indirectly through the impacts of massive particles, via , , or that produces measurable signals like or track patterns in detectors. Electromagnetic waves, however, are detected through their oscillating electric and magnetic fields, exploiting phenomena such as the or resonant absorption in antennas. Additionally, the wavelength relations underscore the disparity; for photons, λ = hc / E, tying wavelength directly to energy without a rest mass term, whereas for matter particles, λ = h / p depends on momentum p = mv, incorporating the particle's mass m. Electromagnetic waves possess transverse polarization arising from their vector field nature, allowing manipulation of oscillation planes, but matter waves lack this electromagnetic polarization, as their wave functions are scalar or spinorial without inherent transverse EM components.

Similarities to Classical Waves

Matter waves display interference and diffraction phenomena that closely parallel those of classical waves, such as the formation of characteristic patterns in Young's double-slit experiment, where electrons or atoms passing through slits produce intensity distributions akin to light or sound waves interfering constructively and destructively. The underpins these behaviors in matter waves, as the linear time-dependent resembles the classical in its form, enabling solutions where multiple waves add linearly to produce effects without altering the fundamental wave propagation. In regions of varying potential, matter waves exhibit and governed by a de Broglie analog of , where the ratio of sines of incidence and refraction angles relates to the square of kinetic energies, mirroring the bending of classical waves at interfaces between media of different speeds. A striking classical analogy to de Broglie matter waves arises in experiments with walking droplets on a vibrating bath, where millimeter-sized droplets—acting as massive "particles"—propagate by bouncing and are guided by the waves they generate on the surface, reproducing and patterns observed in quantum matter waves. Furthermore, wave packets representing localized matter waves, formed by superposing waves of slightly different frequencies and momenta, undergo and similar to classical wave packets, including the production of beats when two such packets overlap, analogous to amplitude-modulated sound waves.

Applications

Electron-Based Technologies

Transmission electron microscopy (TEM) leverages the wave nature of electrons to achieve atomic-scale imaging through interference and diffraction patterns formed as the electron beam passes through thin specimens. In TEM, electrons with de Broglie wavelengths on the order of picometers interact with the sample, producing transmitted beams that interfere to form high-contrast images of atomic structures. This technique, pioneered by Ernst Ruska in the 1930s, enables resolutions approaching 0.05 nm, far surpassing optical microscopy due to the short wavelength of accelerated electrons. Ruska's foundational work on electron optics earned him half of the 1986 Nobel Prize in Physics, recognizing the electron microscope's role in visualizing matter at unprecedented scales. Advancements in aberration-corrected TEM during the and have pushed resolutions to sub-Ångström levels (below 0.1 nm), correcting spherical and chromatic aberrations to preserve phase information in the wave. For instance, a 2002 implementation of computer-controlled aberration correction in (a variant of TEM) demonstrated sub-Ångström by optimizing probe focus, allowing direct imaging of individual atomic columns in crystals. These corrections enhance signal-to-noise ratios and enable of atomic positions and bonding, critical for applications like defect characterization. Scanning electron microscopy (SEM) utilizes focused beams to probe surface , where the diffraction limit imposed by electron wavelengths theoretically supports resolutions down to nanometers, though practical limits arise from beam-sample interactions. In SEM, raster-scanning the beam excites from the surface, and the wave underlying beam formation ensures precise probing of features as small as 1 in modern instruments. This diffraction-limited approach has revolutionized surface analysis in fields like and , providing three-dimensional-like images without requiring vacuum-compatible samples as in TEM. Electron holography extends these principles by recording patterns between a reference electron wave and the wave modulated by the specimen, enabling reconstruction of both and information. Developed from Dennis Gabor's 1948 concept to improve TEM resolution, off-axis electron holography uses a biprism to split and recombine beams, allowing phase shifts—indicative of electrostatic potentials or —to be quantitatively mapped at atomic resolution. This technique is particularly valuable for visualizing invisible phenomena, such as in semiconductors or walls in . Cathode ray tubes (CRTs) and electron accelerators serve as primary sources of coherent electron matter waves for these technologies, accelerating electrons to energies where their de Broglie wavelengths become comparable to atomic spacings. In CRTs, thermionic cathodes emit electrons accelerated by high voltages (typically 10-30 kV), producing beams with wavelengths around 0.01 nm suitable for display and early oscilloscope applications, though primarily operated in the ray optics regime. High-energy electron accelerators, such as linear accelerators, generate relativistic electron waves with even shorter wavelengths (sub-picometer), enabling diffraction-based experiments and feeding advanced microscopes for sub-atomic probing. These sources underpin electron diffraction as the fundamental principle for wave-based imaging in all such devices.

Neutron Scattering Techniques

Neutron scattering techniques exploit the wave nature of , with de Broglie wavelengths on the order of angstroms, to probe and molecular structures in materials and biological samples. Unlike charged particles, interact primarily via the strong nuclear force, enabling deep penetration into bulk materials without significant surface effects. Elastic neutron scattering, a form of neutron diffraction, determines crystal structures by measuring the interference patterns of scattered neutrons with the same energy as the incident beam. This technique reveals atomic positions, including light elements like hydrogen that are challenging for X-ray methods, and has been applied to biological macromolecules such as proteins and DNA hydration networks. For instance, neutron diffraction has elucidated water molecule arrangements around DNA, identifying ordered hydrogen-bonded layers that stabilize the double helix. In materials science, it maps lattice parameters and defects in crystals like metals and semiconductors. Inelastic neutron scattering measures energy transfers between neutrons and the sample, providing insights into dynamic excitations such as —quantized lattice vibrations—and —quantized spin waves in magnetic materials. This has characterized phonon dispersion relations in semiconductors and magnon spectra in ferromagnets, revealing interactions that influence thermal and magnetic properties. For example, inelastic scattering studies have quantified phonon lifetimes in , aiding models of in nanostructures. Small-angle neutron scattering (SANS) probes larger-scale structures, from 1 to 100 nanometers, by detecting neutrons scattered at low angles, ideal for investigating nanostructures like polymers, colloids, and biological assemblies. In , SANS determines the overall shapes and aggregation states of proteins in solution, while in , it analyzes distributions in composites. Contrast variation using labeling enhances resolution for specific components, such as lipid bilayers in membranes. Modern facilities like the , operational since 2006 at , generate intense pulsed neutron beams via proton on a heavy metal target, enabling high-flux experiments for time-resolved studies. Neutrons' high penetration—up to centimeters in metals compared to micrometers for X-rays—allows non-destructive analysis of thick or complex samples, such as geological cores or hydrated biological tissues. Polarized neutron scattering uses neutrons with aligned spins to separate nuclear and magnetic contributions, enabling detailed studies of magnetic structures and domain formations. This has mapped antiferromagnetic ordering in oxides and spin correlations in high-temperature superconductors, providing quantitative measures of magnetic moments. In combination with other techniques, it distinguishes chiral magnetic textures in multiferroic materials.

Atom Interferometry and Sensing

Atom interferometry leverages the wave nature of atoms to achieve unprecedented precision in measuring inertial forces, making it a for advanced sensing technologies. In light-pulse atom interferometers, atoms are manipulated using pulses that act as beam splitters, mirrors, and recombiners, creating matter-wave patterns sensitive to shifts induced by accelerations, rotations, and gravitational fields. The Ramsey-Bordé method, an extension of the original Ramsey separated oscillatory fields technique, employs a sequence of π/2 and π pulses to generate closed interferometric paths, enabling robust measurements even in the presence of Doppler shifts and enabling applications in precision metrology. This configuration has been instrumental in suppressing systematic errors, such as those from , allowing for high-contrast fringes and long interrogation times. One of the primary applications is in , where atom interferometers serve as absolute quantum gravimeters capable of detecting minute variations in the g. These devices have achieved sensitivities below $10^{-9} \, g (equivalent to 1 μGal), with transportable systems demonstrating long-term under 10 nm/s², far surpassing classical instruments for geophysical surveys and resource exploration. For instance, cold-atom gravimeters using cesium or have been deployed in mobile configurations to map anomalies with sub-milligal precision over extended field operations. In sensing, atom-interferometer gyroscopes exploit the Sagnac shift in counter-propagating atomic , providing below 10^{-7} /s and inertial without external references, particularly advantageous in GPS-denied environments. The foundations of atomic timekeeping trace back to the first practical cesium in , which utilized Norman Ramsey's separated oscillatory fields method to achieve frequency stability defining the SI second. Modern cold-atom interferometers emerged in the early , with the first demonstrations using stimulated Raman transitions to create light-pulse interferometers for inertial measurements, paving the way for compact systems in and fundamental physics tests. Cold atom clouds, produced via and magneto-optical trapping, serve as ideal point sources for these interferometers, offering low-velocity spreads and high phase coherence essential for extended baselines. fountain clocks, which launch cooled cesium atoms upward in a symmetric trajectory, extend interrogation times to about 1 second, yielding fractional frequency instabilities below $10^{-15} at 1 second, and form the basis for primary frequency standards at institutions like NIST. Beyond terrestrial applications, atom interferometry has enabled stringent , including measurements of the and violations with sensitivities approaching parts in $10^{15}. Dual-species interferometers using isotopes have constrained the weak to $10^{-8}, providing laboratory-scale probes of curvature. In the 2020s, proposals for space-based atom interferometers, such as those aboard the or dedicated missions like China's cold atom gyroscope satellite, aim to extend these capabilities to microgravity environments, targeting detection analogs and enhanced inertial sensing for deep-space navigation with projected sensitivities improved by orders of magnitude over ground-based systems.

Emerging Molecular and Quantum Applications

Molecular enables quantum simulations of dynamics by leveraging the de Broglie wave nature of molecules to probe effects along different pathways. In such setups, molecular wave packets are split and recombined, analogous to optical interferometers, allowing researchers to observe and manipulate quantum that influences outcomes. For instance, in the study of the H + HD → H2 + D , a phase shift in the molecular wavefunction, induced by a , can steer the reaction toward specific products by constructive or destructive between pathways. This approach provides insights into quantum control of complex chemical processes, where classical simulations fail due to exponential scaling with system size. Optomechanical systems incorporating molecular matter waves offer enhanced sensitivity for force detection at the quantum limit, utilizing the coupling between molecular de Broglie waves and optical cavities to sense ultraweak interactions. These systems exploit the mechanical motion of trapped molecules, whose wave-like behavior amplifies displacement signals through . A notable example is the use of nano-optomechanical resonators to detect forces from single molecules, achieving sensitivities down to zeptonewtons by monitoring cavity frequency shifts induced by molecular vibrations or external fields. Such configurations extend matter-wave to measure intermolecular forces, providing a platform for precision sensing in and . Matter-wave quantum computing with neutral molecules in ion traps represents a frontier in scalable architectures, where molecular superpositions serve as s manipulated via sympathetic cooling and coherent control in the . Prototypes leverage ion traps to indirectly stabilize neutral molecules through interactions, enabling gate operations on molecular rovibrational states with long coherence times. Early demonstrations in the early have achieved entangled states of diatomic molecules like NaK, paving the way for quantum simulation of molecular Hamiltonians beyond classical capabilities. These systems combine the rich internal of molecules with the precision of ion-trap technology for fault-tolerant computing. A landmark 2022 experiment demonstrated quantum superposition of diatomic fermionic NaK molecules, maintained below the Fermi temperature for quantum degeneracy. This achievement, involving evaporative cooling in optical traps, allows testing of predictions, such as many-body correlations and in molecular gases. The superposition enables probes of field-theoretic effects like pair production analogs in curved molecular potentials, bridging and . Hybrid Bose-Einstein condensate (BEC) systems with molecules facilitate analog gravity simulations, particularly for , by engineering sonic horizons in ultracold molecular gases. These setups combine molecular BECs with atomic components to create tunable effective spacetimes, where excitations mimic emission at horizons. Such hybrid platforms simulate gravitational phenomena, including entanglement across horizons, using controllable molecular wave .

References

  1. [1]
    De Broglie's Matter Waves – University Physics Volume 3
    De Broglie's hypothesis of matter waves postulates that any particle of matter that has linear momentum is also a wave. The wavelength of a matter wave ...
  2. [2]
    Matter as a Wave – Physics 132: What is an Electron? What is Light?
    Electrons were the first particles with mass to be directly confirmed to have the wavelength proposed by de Broglie. Subsequently, protons, helium nuclei, ...
  3. [3]
    Quantum interference of large organic molecules - Nature
    Apr 5, 2011 · Although matter wave interference of BECs relies on the de Broglie wavelength of the individual atoms, our massive molecules always appear ...
  4. [4]
    [PDF] Louis de Broglie - Nobel Lecture
    The energy of each quantum has, moreover, a value proportional to the frequency v of the ra- diation. It is equal to hv, h being a universal constant since ...Missing: citation | Show results with:citation
  5. [5]
    [PDF] On the Theory of Quanta Louis-Victor de Broglie (1892-1987)
    Mar 7, 2010 · his thesis (p. 351): “The reason that MAUPERTUIS' integral equals an integer time h, is that each integral is relative to each variable and ...
  6. [6]
    Recherches sur la théorie des Quanta | Annales de Physique
    Ann. Phys., Vol. 10, N°3 (1925), pp. 22–128. Recherches sur la théorie des Quanta. Louis De Broglie. Cet article ne possède pas de résumé.
  7. [7]
    Erwin Schrödinger's Path to Wave Mechanics
    In 1926, Schrödinger invoked Hamilton's optical-mechanical analogy that related ray optics and particle mechanics: His invention of wave mechanics.
  8. [8]
  9. [9]
    [PDF] Clinton J. Davisson - Nobel Lecture
    The de Broglie relation was tested by computing wavelengths from the angles of the diffraction beams and the known constant of the crystal, and comparing these ...
  10. [10]
    Diffraction of Cathode Rays by a Thin Film - Nature
    A fine beam of homogeneous cathode rays is sent nearly normally through a thin celluloid film (of the order 3 × 10 6 cm. thick) and then received on a ...Missing: interference | Show results with:interference
  11. [11]
    Experiments on the diffraction of cathode rays - Journals
    M. L. de Broglie has introduced a theory of mechanics according to which a moving particle behaves as a group of waves whose velocity and wave-length are ...
  12. [12]
    The Nobel Prize in Physics 1937 - NobelPrize.org
    The Nobel Prize in Physics 1937 was awarded jointly to Clinton Joseph Davisson and George Paget Thomson "for their experimental discovery of the diffraction ...
  13. [13]
    Otto Stern and Wave-Particle Duality | SpringerLink
    Jun 20, 2021 · Otto Stern provided the first demonstration for atoms, showing wave properties of helium atoms in 1929, confirming de Broglie's wave-particle ...
  14. [14]
    (IUCr) The early development of neutron diffraction
    These first diffraction experiments were aimed at demonstrating properties of neutrons themselves, namely the wave–particle duality of the neutron. Chadwick's ...
  15. [15]
    [PDF] XXXV. A Tentative Theory of Light Quanta. By LOUIS DE BROGLIE
    On the theoretical side Bohr's theory, which is supported by so many experimental proofs, is grounded on the postulate that atoms can only emit or absorb.
  16. [16]
  17. [17]
    6.5 De Broglie's Matter Waves - University Physics Volume 3
    Sep 29, 2016 · Example 6.11​​ Calculate the de Broglie wavelength of: (a) a 0.65-kg basketball thrown at a speed of 10 m/s, (b) a nonrelativistic electron with ...
  18. [18]
    [PDF] Wolfgang Ketterle - The Magic of Matter Waves - MIT Physics
    However, the wavelength of the de Broglie waves is so small that the wave character of matter only becomes manifest in the microscopic domain. For example ...
  19. [19]
    [PDF] 1.3 THE PHYSICAL CONTENT OF QUANTUM KINEMATICS AND ...
    One finds how to base the probability picture on the theory of the transformation of matrices in the following papers: W. Heisenberg (Zeits. f. Physik, 40, 501 ...
  20. [20]
    Quantisierung als Eigenwertproblem - Wiley Online Library
    Schrödinger. Zürich, Physikalisches Institut der Universität. Search for more papers by this author. First published: 1926. https://doi.org/10.1002/andp.
  21. [21]
    A common assumption in the different theories of Louis de Broglie ...
    May 1, 2021 · In this paper, we present a reconstruction of how group velocity became essential for both de Broglie and Schrödinger, as well as a comparison ...Missing: seminal | Show results with:seminal
  22. [22]
    [PDF] Phase and Group Velocity of Matter Waves—CE Mungan, Spring 2017
    Phase and Group Velocity of Matter Waves—C.E. Mungan, Spring 2017 ... de Broglie relations as E = !ω = hf and p = !k = h /λ in terms of Planck's ...<|control11|><|separator|>
  23. [23]
    Relativistic Derivations of de Broglie and Planck-Einstein Equations
    Aug 1, 2012 · We show how Special Relativity sets tight constraints on the form of possible relations that may exist between the four-momentum of a particle ...
  24. [24]
  25. [25]
  26. [26]
    [PDF] On the quantum mechanics of collision processes.
    “Zur Quantenmechanik der Stoßvorgänge,” Zeit. Phys. 37 (1926), 863-867. On the quantum mechanics of collision processes. [Preliminary announcement (. 1. )] By ...Missing: original | Show results with:original
  27. [27]
    Zur Quantentheorie des Atomkernes
    (Eingegangen am 2. August 1928.) Es wird der Versuch gemacht, die Prozesse der a-Ausstrahlung auf Grund der. Wellenmeehanik n~her zu untersuchen und ...Missing: Kernzerfalls | Show results with:Kernzerfalls
  28. [28]
    Significance of Electromagnetic Potentials in the Quantum Theory
    Mar 3, 2025 · Aharonov and Bohm proposed a scenario in which quantum particles experience electromagnetic effects even though there is no field in their ...
  29. [29]
    The 2001 Nobel Prize in Physics - Popular information
    This year's Nobel Prize in Physics deals with an extreme state of matter, the Bose-Einstein Condensate. The three scientists who are awarded the Prize ...
  30. [30]
    Bose-Einstein Condensate: A New Form of Matter | NIST
    Oct 9, 2001 · ... 1995 with the creation of the world's first Bose-Einstein condensate—a new form of matter. The work earned them, along with Wolfgang ...
  31. [31]
    June 5, 1995: First Bose Einstein Condensate
    Jun 1, 2004 · The world's first BEC was achieved at 10:54 AM on June 5, 1995 in a laboratory at JILA, a joint institute of University of Colorado, Boulder, and NIST.
  32. [32]
    The Nobel Prize in Physics 2001 - NobelPrize.org
    The Nobel Prize in Physics 2001 was awarded jointly to Eric A. Cornell, Wolfgang Ketterle and Carl E. Wieman for the achievement of Bose-Einstein condensation.
  33. [33]
    The Discovery of Superfluidity | Journal of Low Temperature Physics
    Jan 11, 2007 · Superfluidity is a remarkable manifestation of quantum mechanics at the macroscopic level. This article describes the history of its discovery.
  34. [34]
    Theory of ultracold atomic Fermi gases | Rev. Mod. Phys.
    Oct 2, 2008 · This review covers the physics of quantum degenerate atomic Fermi gases, focusing on interactions, superfluid phases, and the BCS, BEC, and ...
  35. [35]
    Quantum Theory of Waves and Particles - University of Pittsburgh
    An essential part of quantum mechanics deals with how matter waves change over time. Mostly, matter waves behave just like ordinary waves. If you have ever ...
  36. [36]
    Stationary States
    The stationary wavefunctions that we have just found are, in essence, standing wave solutions to Schrödinger's equation.<|control11|><|separator|>
  37. [37]
    [PDF] 4. The Infinite Square Well - Physics
    An infinite square well is a model where a particle is confined to a region with zero potential energy inside and infinite potential energy outside.
  38. [38]
    The Quantum Mechanical Model of the Atom - EdTech Books - BYU
    Electron standing waves: Only integer multiples of half-wavelengths can form stable standing waves. The individual electrons within the atom are best ...
  39. [39]
    [PDF] ELECTRON TRANSPORT IN QUANTUM DOTS.
    The deBroglie wavelength of these electrons is comparable to the size of the dot, and the electrons occupy discrete quantum levels (akin to atomic orbitals in ...
  40. [40]
    [PDF] Franck-Hertz experiment, Bohr atom, de Broglie waves
    Oct 7, 2013 · – Standing waves have quantized frequencies, might be related to quantized energies. – Einstein had shown that light, typically thought of as ...
  41. [41]
    Evanescent-mode propagation and quantum tunneling | Phys. Rev. E
    Jul 1, 1993 · We compare experimental data of an electromagnetic wave packet traversing an evanescent waveguide region with theoretical values derived for particle tunneling.
  42. [42]
    Low-energy electron diffraction for surface structure analysis
    Low-energy electron diffraction (LEED) is the most important technique for studying the atomic structures of crystalline surfaces.
  43. [43]
    Low-Energy Electron Diffraction: Experiment, Theory and Surface ...
    Low-Energy Electron Diffraction or LEED has become the prime tech nique used to determine atomic locations at surfaces.
  44. [44]
    Young's Interference of Electrons in Field Emission Patterns
    Jan 4, 2002 · Abstract. We have observed interference fringes of electrons in field emission patterns from multiwalled carbon nanotubes at 60 K. The observed ...Missing: paper | Show results with:paper
  45. [45]
    wavelength of electron | Glossary | JEOL Ltd.
    Wavelength of an electron is calculated for a given energy (accelerating voltage) by using the de Broglie relation between the momentum p and the wavelength λ ...
  46. [46]
    The Young-Feynman controlled double-slit electron interference ...
    Jul 18, 2019 · The Young-Feynman experiment consists of three parts. The first part involves the observation of interference fringes in a double slit setup.
  47. [47]
    [PDF] DETERMINATION OF SCATTERING LENGTHS AND MAGNETIC ...
    At the end of 1973 we put the first powerful neutron interferometer into ope- ration, using dynamical diffraction in a suitable shaped perfect silicon crystal¹² ...
  48. [48]
    Observation of Gravitationally Induced Quantum Interference
    Jun 9, 1975 · We have used a neutron interferometer to observe the quantum-mechanical phase shift of neutrons caused by their interaction with Earth's gravitational field.Missing: Zeilinger Rauch
  49. [49]
    [PDF] NEUTRON INTERFEROMETRY
    3.3 Phase Shift by Gravity and by Interferometer Acceleration. Due to the mass of the neutron a phase shift arises if both beams travel at different ...
  50. [50]
    Ultra-Cold Neutrons in qBounce Experiments as Laboratory for Test ...
    This paper analyzes the effect of chameleon fields on the quantum gravitational states of ultra-cold neutrons (UCNs) in qBounce experiments with mirrors.
  51. [51]
    Search for Dark Energy with Neutron Interferometry - Oxford Academic
    Feb 15, 2024 · It should be noted that ultra-cold neutrons provide versatile tools in the search for new physics in addition to scalar fields with screening ...
  52. [52]
    Atomic interferometry using stimulated Raman transitions
    Jul 8, 1991 · The mechanical effects of stimulated Raman transitions on atoms have been used to demonstrate a matter-wave interferometer with laser-cooled sodium atoms.
  53. [53]
    First-order spatial coherence measurements in a thermalized two ...
    From this, we extract the corresponding thermal de Broglie wavelength λ th = 1.48(1) µm. This agrees well with theoretical value of 1.482(2) µm, obtained by ...
  54. [54]
    The Nobel Prize in Physics 1997 - NobelPrize.org
    The Nobel Prize in Physics 1997 was awarded jointly to Steven Chu, Claude Cohen-Tannoudji and William D. Phillips for development of methods to cool and trap ...
  55. [55]
    Optics and interferometry with atoms and molecules | Rev. Mod. Phys.
    Jul 28, 2009 · Interference with atomic and molecular matter waves is a rich branch of atomic physics and quantum optics. It started with atom diffraction ...
  56. [56]
    Matter-wave interferometers with trapped strongly interacting ...
    May 28, 2024 · We implement two types of matter-wave interferometers using trapped Bose-condensed Feshbach molecules, from weak to strong interactions.Missing: 2020s | Show results with:2020s
  57. [57]
    [PDF] CHAPTER 5 - Portland State University
    A medium is called nondispersive when the phase velocity is the same for all frequencies and equal to the group velocity. All matter waves are dispersing – they ...Missing: seminal | Show results with:seminal
  58. [58]
    [PDF] 9 Particles as waves - Physics Courses
    We can see that the group velocity of the wave coincides with the velocity of the particle itself p/m, however the phase velocity is different. This is another ...
  59. [59]
    [PDF] De Broglie's postulate - wave-like properties of particles
    These postulates introduce a quantization of electromagnetic field: electromagnetic field is emitted, absorbed and transmitted in lumps of energy, called ...Missing: differences | Show results with:differences
  60. [60]
    12.5 The Wave Nature of Matter | Texas Gateway
    The wave nature of matter allows it to exhibit all the characteristics of other, more familiar, waves. Diffraction gratings, for example, produce diffraction ...
  61. [61]
    1.7: Matter as Waves - Chemistry LibreTexts
    Aug 13, 2021 · The idea that electrons can behave as waves, creating interference patterns normally associated with light, is now well-established.
  62. [62]
    [PDF] 4 The Schrödinger wave equation
    De Broglie suggested a relationship between the effective wavelength of the wave function ... The Schrödinger is nominally similar to the classical wave equation.
  63. [63]
    Wave mechanics, from Louis de Broglie to Schrödinger: a comparison
    This paper analyses Schrödinger's derivations of the wave equation, showing the differences and similarities between his theory and de Broglie's. It will be ...
  64. [64]
    De Broglie Wave's Snell's law, Fresnel equations, and Brewster's ...
    We present the quantum reflection rule, Snell's law, and Fresnel equations. We define the concept of transition energy for the de Broglie wave.
  65. [65]
    When fluid dynamics mimic quantum mechanics | MIT News
    such as electrons, or the photons in a beam of light — are borne along on waves of some type, like ...Missing: massive | Show results with:massive
  66. [66]
    Famous Experiment Dooms Alternative to Quantum Weirdness
    Oct 11, 2018 · In describing individual particles guided by corresponding pilot waves, de Broglie didn't account for the way multiple interacting particles ...Missing: massive | Show results with:massive
  67. [67]
    The Theory of Beats and the de Broglie Relation - SCEPHIR
    Jul 16, 2025 · Abstract. The de Broglie relation that associates each elementary particle to a matter wave, the wavelength of which is equal to λ = h / p ...
  68. [68]
    Ernst Ruska – Nobel Lecture - NobelPrize.org
    Award ceremony speech. Nobel Lecture, December 8, 1986. The Development of the Electron Microscope and of Electron Microscopy.
  69. [69]
    The Nobel Prize in Physics 1986 - NobelPrize.org
    The Nobel Prize in Physics 1986 was divided, one half awarded to Ernst Ruska "for his fundamental work in electron optics, and for the design of the first ...
  70. [70]
    Ernst Ruska – Facts - NobelPrize.org
    Ernst Ruska discovered that a magnetic coil could be used as a lens for electron beams and developed the first electron microscope in 1933.
  71. [71]
    Achieving sub-0.5-angstrom–resolution ptychography in ... - Science
    Feb 22, 2024 · For more than two decades, aberration correctors have enabled subangstrom resolution in transmission electron microscopes (TEMs) and scanning ...
  72. [72]
    Scanning Electron Microscopy - Nanoscience Instruments
    Scanning electron microscopy (SEM) uses a focused electron beam to scan a sample's surface, generating high-resolution images and detailed surface information.
  73. [73]
    Progress in atomic-resolution aberration corrected conventional ...
    Aberration-corrected transmission electron microscopes operating at electron energies of typically 200 to 300 keV today offer a Rayleigh resolution of about 50 ...
  74. [74]
    Phase Reconstruction of Low-Energy Electron Holograms of ...
    Nov 11, 2022 · We show that phase imaging can map the projected atomic density of the molecule given by the number of atoms in the electron path.Iterative Reconstruction... · Results · Discussion · Methods
  75. [75]
    How Particle Accelerators Work | Department of Energy
    Jun 18, 2014 · A particle accelerator is a machine that accelerates elementary particles, such as electrons or protons, to very high energies.
  76. [76]
    Accelerators and Nobel Laureates - NobelPrize.org
    Aug 28, 2001 · The most common accelerator today is the cathode ray tube which is used in TV sets and in computer displays. In the tube, a beam of electrons, ...
  77. [77]
    [PDF] An Introduction to Neutron Scattering
    Elastic scattering and definition of the structure factor, S(Q). 4. Coherent & incoherent scattering. 5. Inelastic scattering. 6. Magnetic scattering. 7 ...
  78. [78]
    What makes neutron scattering unique | ORNL
    Apr 1, 2020 · Perhaps an even more important upside to this weak scattering is that neutrons easily penetrate materials that would stop X-rays or electrons, ...
  79. [79]
    Elastic Scattering / MLZ
    Elastic neutron scattering is used to investigate the atomic and magnetic structure of materials, depending on the interaction of the neutrons with the nuclei ...
  80. [80]
    Neutrons take a deep dive into water networks surrounding DNA
    Nov 4, 2021 · Water coats DNA via hydrogen bonds, influencing its shape. Neutrons show water's hydrogen bond patterns, revealing 64 water molecules' ...
  81. [81]
    [PDF] Elastic neutron scattering
    Example: ferromagnetic structure ... The magnetic structure does not necessarily have the same periodicity and symmetry as the underlying crystal structure.
  82. [82]
    [PDF] Introduction to Inelastic Neutron Scattering
    1) Inelastic neutron scattering: ideal probe of phonons and magnetic excitations of all kinds. 2) Different techniques: (variations of triple-axis and time ...
  83. [83]
    Inelastic neutron scattering measurements - Delaire Research Group
    The Delaire group uses inelastic neutron scattering (INS) to measure phonons and magnons in crystalline materials, typically as a function of temperature.
  84. [84]
    Small Angle Neutron Scattering (SANS) | NIST
    Apr 25, 2017 · Small-Angle Neutron Scattering (SANS) probes material structure on the nanometer (10-9 m) to micrometer (10-6 m) scale.
  85. [85]
    Small-Angle Neutron Scattering (SANS) - ScienceDirect.com
    Small-angle neutron scattering is a noninvasive method in which contrast variation and selective labeling are used to extract the spatial arrangement of ...
  86. [86]
    Oak Ridge National Laboratory's Spallation Neutron Source ...
    Apr 28, 2021 · On this day, at 2:04 p.m. on April 28, 2006, the newly constructed Spallation Neutron Source (SNS) delivered its first beam to target.
  87. [87]
    [PDF] The power of polarised neutrons
    Polarised neutrons, with parallel spins, are used to study magnetic materials. Comparing scattering patterns with opposite polarizations helps map magnetic ...
  88. [88]
    Polarized-neutron-scattering studies on the chiral magnetism in ...
    Neutron diffraction with spherical polarization analysis is a powerful tool for studying the multiferroic materials where the ferroelectric polarization arises ...
  89. [89]
    Quantum interference in chemical reactions - Physics Today
    Feb 1, 2018 · The reaction paths on the surface are analogous to particle trajectories passing through slits cut into a screen. Interference is observed ...
  90. [90]
    Ultrasensitive nano-optomechanical force sensor operated ... - Nature
    Jul 5, 2021 · This enabled mechanical detection of ultraweak interactions such as the force exerted by single electronic spins or molecules and investigations ...
  91. [91]
    A molecular optomechanics approach reveals functional relevance ...
    Jun 21, 2023 · Therefore, we and others have developed force sensing techniques to visualize and quantify where, when, and to what extent distinct molecules ...
  92. [92]
    Evaporation of microwave-shielded polar molecules to quantum ...
    Jul 27, 2022 · Here we demonstrate evaporative cooling of a three-dimensional gas of fermionic sodium–potassium molecules to well below the Fermi temperature using microwave ...
  93. [93]
    Ultracold Molecular Collisions: Quasiclassical, Semiclassical, and ...
    Jun 25, 2025 · Ultracold atoms and molecules provide an unprecedented view into the role of quantum effects in chemical reactions.<|separator|>
  94. [94]
    Quantum simulation of Hawking radiation and curved spacetime ...
    Jun 5, 2023 · The first realization of spontaneous Hawking radiation in an analog experiment was in BEC system. ... In addition to pure analog experiments, ...
  95. [95]
    Ramp-up of Hawking Radiation in Bose-Einstein-Condensate ...
    We show the emergence of analog Hawking radiation out of a “quantum atmosphere” region significantly displaced from the horizon. This is quantitatively studied ...Missing: hybrid molecular