Fact-checked by Grok 2 weeks ago

Inelastic scattering

Inelastic scattering is a fundamental process in physics in which an incident particle or wave interacts with a , leading to an exchange of that alters the internal quantum states of the —such as , vibrational, or excitations—while conserving total but not the of the scattered particle. Unlike , where only the direction of motion changes and is preserved, inelastic scattering involves irreversible energy transfer, often resulting in the emission of secondary particles or photons with shifted wavelengths or energies. This phenomenon occurs across various scales, from interactions to high-energy particle collisions, and is characterized by a that the incident particle must exceed to initiate the . Key characteristics of inelastic scattering include its dependence on the incident particle's , , and the target's , often quantified through cross-sections that decrease with increasing in low-energy regimes. In quantum mechanical treatments, it is described using amplitudes modified by absorption parameters (e.g., S_l = \eta_l e^{i 2 \delta_l} with \eta_l < 1), leading to inelastic cross-sections such as \sigma_{\text{inelastic}} = \frac{\pi}{k^2} \sum (2l+1) (1 - \eta_l^2). The process can be coherent or incoherent, with the former preserving phase relationships and the latter leading to diffuse patterns, and it plays a crucial role in probing material dynamics via the dynamical factor S(\mathbf{q}, \omega), which relates intensity to and exchanges. Inelastic scattering manifests in diverse forms, including (inelastic light scattering involving vibrational modes, with Stokes and anti-Stokes shifts), (probing phonons and magnetic excitations in condensed matter), and (high-energy electron-proton collisions revealing substructure like quarks). Applications span multiple fields: in materials science, techniques like electron energy loss spectroscopy (EELS) enable nanoscale chemical mapping; in nuclear physics, it facilitates studies of excited states and reaction mechanisms; and in particle physics, deep inelastic scattering confirmed the quark model of protons, earning the 1990 Nobel Prize in Physics. These methods, often employing time-of-flight or triple-axis spectrometers, provide insights into energy landscapes and collective excitations in complex systems.

General Principles

Definition and Key Characteristics

Inelastic scattering refers to a fundamental interaction process in physics where an incident particle or wave, such as an electron, photon, or neutron, collides with a target system—typically an atom, molecule, or larger medium—resulting in a transfer of energy that excites or alters the internal energy states of the target, such as electronic, vibrational, rotational, or nuclear levels. Unlike elastic scattering, the total kinetic energy of the incident and scattered particles is not conserved, as a portion of the energy (denoted as ΔE) is absorbed by the target to populate these excited states; however, the overall conservation of total energy and linear momentum remains strictly upheld. This process is prevalent across diverse media, including gases, liquids, and solids, and serves as a probe for probing dynamic material properties. Key characteristics of inelastic scattering include the incident particle experiencing a measurable energy loss (or occasionally gain, in cases like ) upon scattering, quantified by the difference between initial energy E_i and final energy E_f < E_i, where the transferred energy \Delta E = E_i - E_f corresponds to discrete quanta matching the target's excitation spectrum. The scattering can be classified as , preserving phase relationships to reveal collective excitations like , or incoherent, averaging over atomic positions to yield information on local dynamics such as . At high incident energies, the probability of inelastic events is generally lower than that of elastic scattering due to the requirement for precise resonance with internal states, though cross-sections increase with energy for processes involving deep excitations. Representative examples encompass the excitation of electrons to higher quantum orbitals, leading to subsequent radiative decay, or vibrational modes in solids that manifest as or annihilation. Energy transfers span a broad range, from millielectronvolts for low-energy to thousands of electronvolts for core-level transitions. A basic schematic of inelastic scattering depicts an incident particle with wavevector \vec{k}_i and energy E_i approaching the target, scattering at an angle with reduced wavevector \vec{k}_f and energy E_f, such that the momentum transfer is \vec{Q} = \vec{k}_i - \vec{k}_f and the energy transfer is \hbar \omega = E_i - E_f > 0, exciting the target to a higher internal state. Historically, inelastic scattering was first experimentally demonstrated in the early through interactions of photons and electrons, with Arthur Compton's 1923 observation of wavelength shifts in scattering off light elements providing foundational evidence for the particle-like nature of light and energy-momentum exchange in collisions.

Comparison to Elastic Scattering

In elastic scattering, both the total kinetic energy and momentum of the system are conserved in the center-of-mass frame, with no or change in the internal quantum states of the interacting particles. This process resembles the collision of ideal balls, where the particles simply redirect without altering their internal structure or energy distribution. In contrast, inelastic scattering involves the transfer of to excite internal of the target or , such as phonons in lattices or transitions in atoms, resulting in a net loss of for the scattered particle. This changes the of the outgoing particle, thereby altering its de Broglie wavelength compared to the incident one. Cross-sections for inelastic processes are generally smaller than those for at energies, as fewer final states are accessible for ; for instance, the of rotational cross-sections to elastic cross-sections in molecular targets decreases with increasing . These differences have notable practical implications in scattering experiments. Inelastic events introduce energy loss and angular spreading, which broaden diffraction patterns in techniques like electron diffraction by convolving the elastic signal with the distribution of inelastically scattered electrons.

Conservation Laws and Energy Transfer

In inelastic scattering, the fundamental conservation laws dictate that the total energy and momentum of the system must remain unchanged before and after the interaction, despite the transfer of energy to internal degrees of freedom of the target. The total energy conservation is expressed as E_{\text{total}} = E_{\text{kinetic, initial}} + E_{\text{internal, initial}} = E_{\text{kinetic, final}} + E_{\text{internal, final}}, where the kinetic energy encompasses both the incident particle and the target, and internal energy includes excitation states such as vibrational, electronic, or nuclear levels. Momentum is conserved vectorially, \vec{p}_{\text{initial}} = \vec{p}_{\text{final}}, ensuring that the change in the incident particle's momentum is balanced by the recoil and any momentum imparted to the target's internal modes. These principles hold in both non-relativistic and relativistic regimes, though in the latter, they are formulated using 4-momentum conservation: p^\mu_{\text{initial}} = p^\mu_{\text{final}}, where p^\mu = (E/c, \vec{p}), accounting for the invariance of the total 4-momentum under Lorentz transformations. Energy transfer in inelastic scattering occurs through \Delta E = E_i - E_f, the difference between the initial and final of the incident particle, which excites the target to higher internal states. This \Delta E can populate discrete energy levels, such as quantized orbitals in atoms (typically on the order of 1–10 ) or vibrational modes in molecules, or continuous spectra, as seen in excitations where high-energy processes access a of states above the particle separation . In quantum terms, the energy quantum transferred is given by \hbar \omega = E_i - E_f = \Delta E, where \hbar \omega represents the energy of the excitation (e.g., a or ), enforced by a function in the cross-section to satisfy . At high energies, relativistic effects become prominent, requiring the use of the Mandelstam variable s = (p_i + p_t)^2 (center-of-mass energy squared) to describe the available energy for transfer, with corrections for mass-energy equivalence ensuring consistency with . The kinematics of inelastic scattering are analyzed most straightforwardly in the center-of-mass (CM) frame, where the total momentum is zero, simplifying the description of momentum transfer \vec{Q} = \vec{k}_i - \vec{k}_f (with \vec{k} as wavevectors) and energy loss. In this frame, the relative motion of the incident particle and target is equivalent to that of a single particle with \mu = \frac{m_1 m_2}{m_1 + m_2} scattering off a fixed center, influencing the allowable scattering angles and the magnitude of \vec{Q}. Inelastic processes require a minimum for the incident particle, determined by the excitation energy of the target; for example, electronic excitations in atoms demand at least ~10 to access the first discrete levels, while processes may require MeV-scale thresholds due to the binding energies involved. This threshold arises from the kinematic constraint that the available in the CM frame must exceed \Delta E, beyond which the differential cross-section opens up for energy transfers to both discrete and continuous states.

Electron Inelastic Scattering

Plasmon and Single-Particle Excitations

In inelastic within solids and gases, two primary types of electronic excitations dominate at low to intermediate losses: collective plasmon modes and individual single-particle transitions. Plasmon excitations arise from the coherent oscillation of the density in metals and semiconductors, treated theoretically as quantized modes in the electron gas model. These modes typically result in discrete energy losses of 10-20 , depending on the material's plasma frequency, and are a hallmark of the response in condensed matter. Such plasmon excitations are routinely observed and characterized using (EELS) in transmission electron microscopes, where the scattered electron spectrum reveals sharp peaks corresponding to bulk or surface . The theoretical foundation for these collective excitations was established in the , which describes the long-range interactions leading to oscillations with dispersion relations that can be probed via momentum transfer in experiments. In contrast, single-particle excitations involve the promotion of an individual electron from the to conduction band, creating an electron-hole pair, with energy losses spanning a continuum above the bandgap. These are particularly prominent in semiconductors and insulators, where the band structure dictates the available final states. The dispersion of single-particle excitations in the energy-momentum plane often manifests as the Bethe ridge, a feature predicted by the for fast electrons interacting with a free-electron-like target, where the maximum loss occurs when the momentum transfer matches the of electron-electron collisions. This ridge shifts with increasing momentum transfer q and provides insight into the electronic structure beyond simple optical limits. The probability of these inelastic processes is quantified by the differential cross-section, which scales roughly with the square of the a_0^2 \approx 0.0028 \, \mathrm{nm}^2 in for interactions. Consequently, the is given by \lambda_\mathrm{inel} \approx 1 / (n \sigma), where n is the target and \sigma the total inelastic cross-section; for typical solids at incident energies exceeding 1 keV, \lambda_\mathrm{inel} ranges from 10-100 nm, making inelastic events less frequent than but still significant for energy dissipation. In practical applications, such as (TEM), inelastic scattering contributes to beam broadening, where multiple events—both elastic and inelastic—spread the incident probe over a larger area within the sample, limiting for thick specimens. This is particularly pronounced for low-energy losses like plasmons, as the scattering angles remain small (\theta \approx \Delta E / (2 E_0), with \Delta E the energy loss and E_0 the incident energy), yet cumulatively degrade image contrast and spectroscopic signals.

Deep Inelastic Scattering on Nucleons

Deep inelastic scattering (DIS) on nucleons involves the interaction of high-energy electrons with protons or neutrons, enabling the probing of their internal at the quark-gluon level. In this process, an incident exchanges a spacelike with a (valence or sea) within the , leading to a detectable scattered and a spray of hadrons from the struck quark's fragmentation. The "deep" regime is defined by large values of the negative squared transfer Q^2 > 1 GeV² and the Bjorken scaling variable x (typically $0 < x < 1), where x approximates the longitudinal momentum fraction of the nucleon's momentum carried by the interacting quark. These conditions ensure the virtual photon's short wavelength resolves the point-like nature of quarks, distinguishing DIS from shallower inelastic processes. The foundational experiments revealing this quark substructure were conducted at the Stanford Linear Accelerator Center (SLAC) during the late 1960s. Led by , , and , these studies measured electron-proton and electron-deuteron scattering cross-sections using beams up to 20 GeV, observing approximate scaling in the extracted structure functions that indicated composite, point-like constituents inside the nucleon. This work provided direct experimental confirmation of quarks as the building blocks of hadrons, aligning with theoretical predictions and revolutionizing particle physics. For their pioneering investigations concerning deep inelastic scattering of electrons on protons and bound neutrons, Friedman, Kendall, and Taylor shared the 1990 Nobel Prize in Physics. To explain the scaling observed in these experiments, Richard P. Feynman proposed the parton model in 1969, envisioning the nucleon as a collection of quasi-free, point-like partons (subsequently identified as and ) that scatter incoherently off the virtual photon at high energies, akin to partons in a fast-moving jet. In this intuitive framework, the structure functions directly reflect the momentum distributions of these partons within the nucleon. The kinematics of DIS are characterized by the squared momentum transfer Q^2 = -q^2 = 4 E_e E_{e'} \sin^2(\theta/2), where E_e and E_{e'} are the initial and final electron energies, respectively, and \theta is the electron scattering angle in the laboratory frame. This quantity, along with the energy transfer \nu = E_e - E_{e'}, defines the probe's resolution and the Bjorken x. The inclusive cross-section is parameterized by nucleon structure functions, notably F_2(x, Q^2), which measures the total quark momentum distribution weighted by electric charges. In the naive parton model, F_2(x) scales as a function of x alone, but quantum chromodynamics (QCD) introduces scaling violations through Q^2-dependent evolution, driven by gluon emissions and quark-antiquark pair production, as captured in the renormalization group equations. These effects allow extraction of parton distribution functions that evolve predictably with energy scale.

Photon Inelastic Scattering

Compton Scattering

Compton scattering describes the inelastic interaction in which an incident photon collides with a free or loosely bound electron, transferring a portion of its energy and momentum to the electron, causing it to recoil while the scattered photon emerges with a longer wavelength. This process exemplifies the quantum mechanical treatment of photons as particles with momentum p = h / \lambda, where the collision resembles a billiard-ball-like exchange governed by conservation of energy and momentum. The energy transfer to the electron can range from negligible at small scattering angles to significant at larger angles, up to a maximum of nearly the full photon energy for backscattering, highlighting the relativistic kinematics involved when photon energies approach or exceed the electron rest mass energy of 511 keV. The phenomenon was experimentally observed and theoretically explained by Arthur H. Compton in 1923, through measurements of X-ray scattering by graphite and other light elements, where the scattered radiation exhibited a wavelength shift independent of the incident X-ray wavelength. Compton's analysis demonstrated that the observed spectral shift could only be accounted for by treating X-rays as discrete quanta interacting with individual electrons, providing key evidence for the particle nature of light and challenging classical wave theories of scattering. For his discovery of this effect, Compton shared the 1927 Nobel Prize in Physics with C. T. R. Wilson. The quantitative relation for the wavelength shift in Compton scattering by a free electron at rest is given by the Compton formula: \Delta \lambda = \lambda' - \lambda = \frac{h}{m_e c} (1 - \cos \theta) where \lambda and \lambda' are the wavelengths of the incident and scattered photons, respectively, h is , m_e is the , c is the , and \theta is the scattering angle of the photon. This shift, known as the \lambda_C = h / (m_e c) \approx 2.426 \times 10^{-12} m, is purely kinematic and arises from the relativistic conservation laws, with no dependence on material properties for free electrons. In materials such as solids, where electrons are bound to atoms, the simple Compton formula is modified due to the initial momentum distribution of the electrons in their orbitals, leading to a Doppler broadening of the scattered photon spectrum described by the Compton profile J(p_z), which represents the projection of the electron momentum density along the scattering direction. This profile provides insights into the electronic structure, as the broadening reflects the electron's momentum spread rather than a delta-function recoil, and high-resolution measurements using synchrotron radiation or gamma-ray sources have been employed to map these distributions in various solids. For loosely bound electrons, such as in outer shells or conduction bands, the free-electron approximation holds closely, but inner-shell electrons exhibit more pronounced deviations due to stronger binding effects.

Raman Scattering

Raman scattering is a form of inelastic photon scattering where an incident photon interacts with a molecule, leading to a scattered photon with a shifted frequency that corresponds to the energy differences between molecular vibrational or rotational states. This phenomenon, first observed by in 1928 during experiments with monochromatic light passing through liquids, results in a small fraction of scattered light—approximately one in a million photons—exhibiting wavelength changes dependent on the sample's molecular structure. The process involves the incident photon temporarily exciting the molecule to a virtual intermediate state, not a real electronic excited state, from which it relaxes while transferring energy to or from the molecule's vibrational or rotational modes./04:_Chemical_Speciation/4.03:_Raman_Spectroscopy) The frequency shift in Raman scattering manifests as Stokes or anti-Stokes lines relative to the incident frequency ν₀. In Stokes scattering, the scattered frequency ν_s is lower than ν₀ (Δν = ν₀ - ν_s > 0, red-shifted), occurring when the molecule transitions from the ground vibrational state to a higher one, absorbing energy from the photon; this is more probable because most molecules occupy the ground state at room temperature. Anti-Stokes scattering features a higher scattered frequency (Δν < 0, blue-shifted), where the molecule starts in an excited vibrational state and relaxes to the ground state, donating energy to the photon; its lower intensity arises from the Boltzmann distribution favoring lower energy states. The magnitude of the shift equals the energy of the vibrational or rotational transition, given by Δν = E_vib / h, where E_vib is the molecular energy level spacing and h is Planck's constant. Unlike () absorption spectroscopy, which requires a change in the molecule's for a transition to be active, is governed by selection rules based on changes in molecular —the ease with which the electron cloud distorts under an . Vibrational modes that alter polarizability, such as symmetric stretches in nonpolar molecules like O₂ or N₂, produce strong Raman signals but are IR-inactive, making complementary to for identifying molecular symmetries and structures. Asymmetric modes or bends, which change the dipole moment, yield weak Raman activity but strong IR absorption. To overcome the inherently weak Raman signal, enhancement techniques like surface-enhanced Raman spectroscopy (SERS) exploit plasmonic effects on nanostructured metal surfaces, such as roughened silver or gold. Discovered in the mid-1970s through observations of intensified pyridine signals on electrochemically roughened silver electrodes by Fleischmann and colleagues in 1974, SERS achieves enhancement factors of 10⁵ to 10⁶ via electromagnetic field amplification from localized surface plasmons. Subsequent theoretical work by Moskovits in 1978 linked this to plasmon resonances, enabling single-molecule detection by the late 1990s, as demonstrated by Nie and Emory in 1997 using silver nanoparticle aggregates with enhancement up to 10¹⁴–10¹⁵ near "hot spots."

Neutron Inelastic Scattering

Nuclear and Atomic Excitations

In neutron inelastic scattering leading to excitations, fast s with energies typically exceeding 1 MeV interact directly with the through (n,n') reactions, wherein the incident excites discrete levels before being re-emitted with reduced . This process contrasts with interactions involving thermal neutrons, which have average energies around 0.025 and primarily form a compound upon capture, with subsequent de-excitation potentially mimicking inelastic outcomes through channels, though such events are rare due to the low incident energy. Overall, energy transfer in these interactions adheres to laws, where the imparts sufficient to overcome and barriers. For atomic excitations, inelastic neutron scattering can involve energy transfers to individual atomic motions, such as molecular rotations or vibrations in gases or dilute systems, revealing local dynamics and interatomic potentials, though such processes are typically weaker and often overlap with collective excitations in condensed matter. The excited nucleus resulting from inelastic scattering de-excites by emitting gamma rays to return to the ground state. For higher excitation energies, additional de-excitation channels such as particle emission may become possible. The probability of these interactions, quantified by the inelastic scattering cross-section \sigma_{inel}, varies significantly with the target nucleus and incident neutron energy; it is generally small for light nuclei but increases for heavier elements like uranium, where thresholds are lower (e.g., ~0.044 MeV for ^{238}U). Cross-sections rise above the reaction threshold and can approach or exceed those of elastic scattering at higher energies, influencing neutron transport in nuclear systems. Inelastic scattering requires a minimum incident energy, known as the E_{th}, to enable while conserving . For a (n,n') exciting the target to energy E_{exc}, the threshold in the laboratory frame is given by E_{th} = E_{exc} \left(1 + \frac{m_n}{M}\right), where m_n is the and M is the target ; this accounts for the endothermic nature of the (Q-value = -E_{exc}). For example, carbon requires E_{th} \approx 4.8 MeV due to its higher first level, while heavier nuclei have lower thresholds. In nuclear reactors, inelastic scattering plays a key role in slowing down fast neutrons, particularly in fast breeder designs like fast breeder reactors (LMFBRs), where it contributes to spectrum hardening and in coolants such as sodium without traditional moderators. This process affects economy, breeding ratios, and shielding requirements, as excited nuclei produce prompt gamma rays that influence dose rates.

Scattering from Collective Modes

In inelastic neutron scattering, collective modes in condensed matter systems, such as and magnons, give rise to distinct and transfers that reveal the underlying spectra. involves the creation or annihilation of quantized lattice vibrations, where the incident transfers an \hbar \omega_{\vec{q}} to a mode characterized by wavevector \vec{q}. This process allows mapping of the \omega(\vec{q}), which describes how the frequency of lattice vibrations varies with , providing insights into interatomic forces and material elasticity. Measurements of these dispersions are typically performed using triple-axis spectrometers, instruments that sequentially select incident and , analyze scattered neutrons, and detect their , enabling precise resolution of the scattering function in reciprocal space. The triple-axis technique was pioneered by in the 1950s at the , where it was first used to observe dispersions in materials like aluminum. Magnon probes spin-wave excitations in magnetic materials, where neutrons interact with spin fluctuations, transferring \hbar \omega_{\vec{q}} and \hbar \vec{q} to create or destroy s. In antiferromagnets, these inelastic signals often appear prominently at zone boundaries in reciprocal space, corresponding to short-wavelength spin waves near the Brillouin zone edge, and their dispersion reflects the exchange interactions between magnetic moments. Early observations of dispersions were achieved using triple-axis spectrometers on materials like , demonstrating the technique's capability to resolve magnetic excitations with energies on the order of meV. Quasielastic scattering from collective modes arises in systems with diffusive , such as diffusing particles or relaxational processes, where the energy transfer is near zero but broadened by Doppler effects from thermal motions. This manifests as a broadening in the scattering profile, with the proportional to the coefficient and inversely related to the correlation time of the motion, typically resolved at energy scales of ~1 meV using high-resolution spectrometers like backscattering instruments. Such broadening encodes information on dynamic correlations in liquids, polymers, and biological systems. The development of inelastic scattering for collective modes began in the 1950s at nuclear reactors, including , and advanced significantly with facilities like the Institut Laue-Langevin (ILL) established in the 1970s, which provided high neutron fluxes for detailed studies. These experiments measure the dynamic S(\vec{q}, \omega), defined as S(\vec{q}, \omega) = \int d^3 r \int_{-\infty}^{\infty} dt \, \langle \rho^*(\vec{0}, 0) \rho(\vec{r}, t) \rangle e^{i (\vec{q} \cdot \vec{r} - \omega t)}, where \rho is the density operator, which directly encodes the space- and time-dependent correlations of atomic or densities responsible for the observed excitations; this formalism was introduced by Léon van Hove in 1954.

Inelastic Scattering in Molecular and Atomic Systems

Vibrational and Rotational Transitions

In inelastic scattering processes within gaseous or clustered molecular systems, a portion of the incident translational is transferred to the internal of the colliding molecules, specifically exciting or de-exciting vibrational and rotational modes. Vibrational transitions typically involve energy changes on the order of 1000 cm⁻¹, corresponding to the spacing between quantized vibrational levels in diatomic or polyatomic molecules, while rotational transitions occur at much lower energies around 10 cm⁻¹, reflecting the smaller spacing in rotational levels. This translational-to-internal energy partitioning is governed by thermal conditions where the available energy ΔE is comparable to (approximately 200 cm⁻¹ at ), allowing redistribution primarily to rotational modes due to their , though vibrational becomes feasible at higher collision energies. Propensity rules in these collisions favor vibration-translation (V-T) energy transfer processes, particularly those with small changes in vibrational (Δv ≈ ±1) and minimal rotational perturbations, as dictated by energy gap and conservation principles. These rules arise from the dynamics of the interaction potential during close encounters, where near-resonant energy matching enhances transition probabilities, though they can break down in systems with strong intermolecular forces like hydrogen bonding. For polyatomic molecules, the inelastic rate constant k_inel is expressed as the product of the collision frequency Z—given by Z = n σ ⟨v⟩, where n is the , σ the collision cross-section, and ⟨v⟩ the relative mean speed—and the inelastic probability P_inel per collision, with P_inel typically ranging from 0.1 to 1 depending on the molecular complexity and temperature. In , such inelastic collisions play a crucial role in relaxing vibrationally excited molecules without inducing chemical reactions, thereby influencing energy balance and trace species lifetimes. For instance, collisions involving vibrationally excited radicals (v ≥ 1) with atmospheric constituents like N₂ or O₂ lead to efficient V-T deactivation, with rate constants on the order of 10⁻¹¹ cm³ molecule⁻¹ s⁻¹, preventing prolonged excited-state persistence that could otherwise drive unwanted . Similarly, O₂ in excited vibrational states relaxes through non-reactive encounters with ground-state partners, maintaining local in the upper atmosphere. Quantum mechanical treatments of these processes employ close-coupling equations to compute state-to-state cross-sections, accounting for the coupled dynamics of translational, rotational, and vibrational coordinates during the collision. These equations, solved numerically on surfaces, yield detailed differential and integral cross-sections for specific initial and final quantum states, such as transitions in CO(v=2) colliding with He atoms, where rotational propensity rules dictate Δj ≈ ±2 preferences. This approach provides rigorous predictions validated against experimental measurements, essential for understanding propensity and energy transfer efficiencies in polyatomic systems.

Atomic Electronic Excitations

In atomic systems, inelastic scattering primarily involves electronic excitations, where collision energy is transferred to promote electrons to higher orbitals, often leading to emission of photons or ionization. For example, in noble gas atom collisions like He + Ne, inelastic processes excite metastable states with energy transfers of several eV, governed by selection rules and quantified by cross sections from quantum defect theory or experimental beam techniques. These differ from molecular cases by lacking vibrational/rotational structure, focusing instead on discrete electronic levels, and play key roles in plasma physics and astrophysical environments.

Quasielastic and Reactive Collisions

Quasielastic scattering refers to collision processes in molecular and systems where the is small, typically on the order of the kT (where k is Boltzmann's and T is ), resulting in a broadening of the peak rather than distinct inelastic features. This broadening manifests as a broad linewidth in the spectrum, arising from motions such as translational or rotational reorientation of molecules, which occur over timescales from picoseconds to nanoseconds. In quasielastic (QENS), for instance, the incoherent scattering from atoms in polymers reveals these , with linewidths increasing with due to enhanced segmental relaxations and above the . Unlike purely elastic scattering, quasielastic processes conserve momentum approximately while allowing minor energy exchanges that probe diffusive behaviors in complex systems like polymer melts or liquids confined in nanopores. State-resolved studies using molecular beams, pioneered in the 1970s, have observed rainbow scattering patterns in quasielastic collisions, where classical trajectory deflections due to anisotropic potentials lead to intensity maxima in differential cross sections, linking experimental angular distributions to potential energy surfaces. Reactive inelastic scattering involves larger energy transfers that surpass simple excitations, enabling bond breaking and chemical rearrangement, as exemplified by the prototypical H + H₂ → H₂ + H exchange . This differs from non-reactive inelastic scattering primarily through the topology of the (PES), where a path crosses a barrier or , facilitating product formation via short-lived collision complexes. At thermal collision energies (~0.04 eV), reactive cross sections can range from ~0.1 to 10 Ų, depending on the system. Van der Waals complexes often mediate these reactions by trapping reactants in quasi-bound states, enhancing reactivity through resonances that appear as oscillations in the integral cross sections, as seen in the F + → HF + D pathway.

Applications

In Particle Physics and Nuclear Structure

In particle physics, inelastic scattering has been instrumental in probing the substructure of nucleons and revealing the existence of fundamental constituents. Deep inelastic electron-proton scattering experiments conducted at the Stanford Linear Accelerator Center (SLAC) from 1968 to 1973 provided the first evidence for point-like partons within the proton, challenging the then-prevailing view of nucleons as fundamental particles. These experiments, led by Jerome Friedman, Henry Kendall, and Richard Taylor, measured the scattering of high-energy electrons off protons and observed a scaling behavior in the cross-sections that aligned with the quark-parton model proposed by James Bjorken and Richard Feynman. The results indicated that protons consist of fractionally charged, spin-1/2 quarks interacting via the strong force, earning the researchers the 1990 Nobel Prize in Physics. Neutrino inelastic scattering complements by directly probing weak interactions through charged-current processes, where a exchanges a with a in the target . At , experiments such as those performed by the collaboration have measured inclusive charged-current on targets at energies around 6 GeV, revealing details of quasi-elastic and deep inelastic contributions to the cross-section. These measurements help validate models of weak interactions and structure, including the role of axial-vector form factors, and provide benchmarks for experiments like . By isolating charged-current events, such disentangles electroweak parameters and tests the standard model's predictions for - couplings. In nuclear structure studies, inelastic scattering reactions, particularly electron-induced processes like (e, e'n), map collective excitations such as giant resonances, which reflect the shell-like organization of nucleons within the nucleus. Experiments at facilities like Jefferson Lab have used high-resolution inelastic electron scattering to excite and resolve giant dipole and quadrupole resonances in nuclei, demonstrating their sensitivity to isovector and isoscalar modes that arise from coherent proton-neutron motions. These resonances, observed between 10-40 MeV excitation energy, provide empirical tests of nuclear shell models and effective interactions, linking microscopic quark-level dynamics to macroscopic nuclear properties. The angular momentum-resolved spectra further constrain deformation parameters and compression modes, aiding in the understanding of nuclear stability and reactions. Modern high-energy colliders continue to leverage inelastic scattering for fundamental discoveries, such as the search for the at the (LHC). Inelastic proton-proton collisions at 7-13 TeV, analyzed by the ATLAS and collaborations, led to the 2012 observation of the through its decays into photons, W/Z bosons, and bottom quarks, confirming the mechanism of electroweak . The dominant production mode, fusion, exemplifies inelastic scattering where quarks and from colliding protons interact to form the Higgs field mediator. Looking ahead, the Electron-Ion Collider (EIC) at , with construction commencing post-2025 following RHIC shutdown, will enable precision electron-ion collisions to map the three-dimensional quark- structure inside protons and nuclei, including saturation effects at small x. This facility promises deeper insights into , such as the emergence of quark- plasma properties, through tagged .

In Condensed Matter and Materials Science

In , inelastic () serves as a powerful probe for mapping dispersions, which reveal the dynamics underlying material properties in solids. By measuring and momentum transfers, INS provides direct access to phonon frequencies and lifetimes, enabling the characterization of vibrational modes that influence thermal conductivity, electron-phonon coupling, and phase transitions. For instance, in iron-based superconductors like BaFe_2As_2, INS experiments combined with lattice dynamical calculations have identified anomalous phonon softening near the Fe-As bond, linking these modes to enhanced electron-phonon interactions that may contribute to . In high-temperature , such as Bi_{1.5}Pb_{0.55}Sr_{1.6}La_{0.4}CuO_{6+\delta}, INS has detected paramagnon excitations—overdamped spin fluctuations—that mediate formation, highlighting the interplay between magnetic and . Electron energy-loss spectroscopy (EELS) integrated with (TEM) extends inelastic scattering to nanoscale electronic structure analysis, particularly for profiling band gaps and defects in . EELS quantifies energy losses from or electrons, allowing spatial resolution down to atomic scales to map local band structures and identify defect-induced states. In semiconductors like transition metal dichalcogenides, EELS has measured band gaps as low as 1.8 eV with sub-nanometer precision, revealing strain or doping effects on electronic properties. For defect characterization in , such as oxygen vacancies in oxides, core-loss EELS detects fine-structure shifts that correlate with charge redistribution, providing insights into transport limitations in devices like solar cells. Recent advancements in monochromated EELS have improved energy resolution to below 10 meV, enabling the study of excitonic effects in two-dimensional materials. The high neutron scattering cross-section of makes INS particularly advantageous for studying dynamics in hydrogenous materials, such as polymers and battery electrolytes, where coherent scattering from light atoms dominates. In polymer systems, quasi-elastic neutron scattering (QENS) probes segmental motions and chain relaxations, elucidating how hydrogen bonding influences mechanical properties. For lithium-ion batteries, INS has tracked Li^+ diffusion in polymer electrolytes like polyethylene oxide (PEO) doped with LiTFSI, revealing nanosecond-scale dynamics that limit ionic at low temperatures. Post-2010 developments, including higher-flux sources, have enabled in operando studies of Li dynamics in solid-state batteries, showing how polymer chain adsorption on cathode particles impedes ion transport. These insights have guided the design of faster-charging electrolytes with enhanced Li mobility. As of 2025, neutron spin-echo (NSE) techniques have advanced the investigation of slow dynamics in quantum materials, accessing timescales up to hundreds of nanoseconds for processes like spin fluctuations in topological insulators. NSE resolves intermediate scattering functions with high energy resolution, distinguishing diffusive from coherent motions in systems where conventional INS lacks sensitivity. In Kitaev materials like \alpha-RuCl_3, NSE measurements have quantified spin gaps and recovery times exceeding 10 ns at magnetic peaks, linking magnetic dynamics to quantum spin liquid properties. Applications to frustrated magnets and quantum spin liquids, as highlighted in recent facilities like the SNS-NSE, demonstrate NSE's role in probing emergent phenomena, such as magnon-phonon hybridization in low-dimensional systems. These capabilities have illuminated slow relaxation in topological phases, informing potential quantum computing architectures.

In Spectroscopy and Imaging Techniques

Inelastic scattering plays a pivotal role in and techniques by enabling the probing of molecular, atomic, and electronic structures without destructive . These methods leverage energy transfers during events to map chemical compositions, electron densities, and vibrational modes, providing high-resolution insights into materials at micro- and nanoscale levels. In and materials analysis, such techniques facilitate label-free diagnostics and non-invasive inspections, with applications ranging from cellular to security screening. Raman , a cornerstone of these techniques, utilizes inelastic light to generate molecular fingerprints, allowing spatial mapping of biochemical variations in biological samples. Since the , it has been applied in for identifying cancer cells by detecting alterations in , proteins, and nucleic acids, achieving sensitivities up to 98% in distinguishing leukemic lines from normal cells. For instance, in studies, Raman spectra reveal HER2 overexpression and correlate spectral changes with tumor aggressiveness, enabling non-invasive and diagnostics with specificities exceeding 90% in gastrointestinal and skin cancers. This approach integrates with multivariate analysis for automated classification, supporting of thousands of cells in minutes. Compton tomography employs inelastic X-ray scattering to reconstruct three-dimensional distributions, offering a non-destructive method for dense objects. In cargo scanning, backscattered photons from Compton interactions provide material discrimination based on , where organic contraband appears brighter due to higher scattering efficiency at energies of 4-9 MeV. This technique supports security applications at ports and borders, enabling rapid detection of explosives or weapons without unpacking containers, as demonstrated in systems like the U.S. Container Security Initiative. Electron energy-loss spectroscopy (EELS) mapping in (STEM) achieves atomic-resolution imaging of elemental and compositional features through . By analyzing energy-loss near-edge structures, it maps oxidation states and defect distributions, such as in oxides where Mn valence variations are resolved at the single-atom column level with sub-angstrom precision. Advances in aberration-corrected STEM and monochromators have enabled resolutions down to 0.073 nm, facilitating vibrational and mapping in materials like and for defect analysis. Recent advances in coherent Raman scattering (CRS) enhance label-free imaging by coherently amplifying Raman signals, achieving hyperspectral acquisition in microseconds over broad bandwidths for real-time chemical selectivity. Techniques like and detect metabolites in live cells at sub-micromolar sensitivities, applied in biomedicine to study lipid dynamics in cancer and neurodegeneration without exogenous labels. Up to 2025, integration of with CRS and traditional Raman methods has enabled real-time spectral unmixing and denoising via , reducing analysis times to seconds and improving accuracy in complex biological datasets through open-source tools and principles.

References

  1. [1]
    Inelastic Scattering - an overview | ScienceDirect Topics
    Inelastic scattering is defined as a process where an incident particle excites an atom to a higher electronic or nuclear state, resulting in the emission ...
  2. [2]
    Elastic and Inelastic Scattering - Richard Fitzpatrick
    ), whereas the inelastic cross-section is inversely proportional to the particle velocity. Consequently, as the incident particle velocity decreases, inelastic ...
  3. [3]
    Inelastic Scattering - an overview | ScienceDirect Topics
    Inelastic scattering is defined as a process where a particle is excited to a virtual state and then nearly simultaneously de-excites to a different vibrational ...
  4. [4]
    Inelastic Scattering | Definition, Types & Example - Study.com
    May 18, 2025 · Inelastic scattering is a process where an incoming particle or wave, such as a photon, electron, or neutron, transfers some of its energy to ...
  5. [5]
    Scattering
    Inelastic scattering can be a complex and interesting process. It was with high energy inelastic scattering of electrons from protons that the quark structure ...<|control11|><|separator|>
  6. [6]
    [PDF] Inelastic Scattering
    May 4, 2021 · The goal of this book is to describe the underlying scattering physics and dynamic processes in materi- als, and show opportunities to elevate ...
  7. [7]
    Key Characteristics of Inelastic Scattering | nuclear-power.com
    Key Characteristics of Inelastic Scattering are summarized here. Inelastic scattering conserves momentum, while kinetic energy does not conserve.
  8. [8]
    A Quantum Theory of the Scattering of X-rays by Light Elements
    Jan 27, 2025 · The hypothesis is suggested that when an X-ray quantum is scattered it spends all of its energy and momentum upon some particular electron.Missing: discovery | Show results with:discovery
  9. [9]
    Neutron Elastic Scattering | Definition | nuclear-power.com
    The elastic scattering conserves both momentum and kinetic energy of the “system”. It may be modeled as a billiard ball collision between a neutron and a ...<|separator|>
  10. [10]
    Delbr\"uck contribution in the elastic scattering of photons
    Mar 31, 2005 · The process is called elastic because no energy is transferred to the internal degrees of freedom of the atom, which remain unchanged, and no ...
  11. [11]
    Internal Excitation, Inelastic Scattering
    When target and/or projectile have internal degrees of freedom beyond the spins discussed in Sect. 2.8, then their internal states can be excited or ...
  12. [12]
    Elastic and inelastic electron scattering cross sections of ...
    This paper measures elastic and inelastic electron scattering cross sections of trichlorofluoromethane, a CFC with a high ozone depletion potential. Results ...<|control11|><|separator|>
  13. [13]
    Inelastic Scattering in Reflection High-Energy Electron Diffraction ...
    Dec 1, 1997 · Diffraction spots broadened due to surface roughness and inelastic scattering can have identical shape. Consequently separation is possible ...
  14. [14]
    [PDF] Relativistic Kinematics II - UF Physics
    Aug 26, 2015 · Relativistic kinematics uses 4-vectors to simplify problems, including particle decay and scattering, and accounts for conservation of energy ...
  15. [15]
    [PDF] Scattering Theory
    (iii) conservation of momentum, (iv) conservation of spin projection, other conservation laws/constraints, if any. Set the volume equal to unity, because it ...
  16. [16]
    A Collective Description of Electron Interactions. I. Magnetic ...
    A new approach to the treatment of the interactions in a collection of electrons is developed, which we call the collective description.Missing: plasmon | Show results with:plasmon
  17. [17]
    [PDF] Cross Sections for Inelastic Scattering of Electrons by Atoms
    For electron energies not negligible compared to mc2 = 511 keV, one must use relativistic kinematic s. The notion of the momentum transfer 11~ may be most.
  18. [18]
    Mean free path of inelastic electron scattering in elemental solids ...
    Mar 5, 2008 · The characteristic measures of the scattering are the cross section σ or mean free path λ . By definition, they are related through the ...
  19. [19]
    Electron-beam broadening in electron microscopy by solving the ...
    Dec 4, 2020 · In this work, we address beam broadening in a more general way. We numerically solve the electron transport equation without any simplifications.Article Text · INTRODUCTION · METHODS · RESULTS AND DISCUSSION
  20. [20]
    [PDF] 18. Structure Functions - Particle Data Group
    Dec 1, 2023 · The double-differential cross section for deep inelastic scattering can be expressed in terms of kinematic variables in several ways. d2σ dx dy.
  21. [21]
    The Nobel Prize in Physics 1990 - NobelPrize.org
    The Nobel Prize in Physics 1990 was awarded jointly to Jerome I. Friedman, Henry W. Kendall and Richard E. Taylor for their pioneering investigations.
  22. [22]
    [PDF] R.P. Feynman, California Institute of Technology, Pasadena
    THE PARTON PICTURE OF DEEP INELASTIC ELECTRON SCA'ITERING. The wave function for a proton moving with a large momentum P to the right along t~e z axis, ...
  23. [23]
    A century of Compton scattering | Nature Reviews Physics
    Mar 28, 2023 · Compton scattering also has practical applications in other fields. It is the dominant scattering process of gamma rays and high-energy X-rays ...
  24. [24]
    An Overview of the Compton Scattering Calculation - MDPI
    The Compton scattering process plays significant roles in atomic and molecular physics, condensed matter physics, nuclear physics and material science.
  25. [25]
    [PDF] A Quantum Theory of the Scattering of X-Rays by Light Elements
    The experimental support of the theory indicates very convinc- ingly that a radiation quantum carries with it directed momentum as well as energy. Emphasis has ...Missing: discovery | Show results with:discovery
  26. [26]
    Arthur H. Compton – Facts - NobelPrize.org
    Arthur Holly Compton, Nobel Prize in Physics 1927. Born: 10 September 1892, Wooster, OH, USA. Died: 15 March 1962, Berkeley, CA, USA.
  27. [27]
    Compton scattering revisited - ScienceDirect.com
    We review the standard theory of Compton scattering from bound electrons, and we describe recent findings that require modification of the usual ...
  28. [28]
    Compton Scattering as a Technique for the Study of Solids
    The advantages and insights afforded by Compton profile measurements in the study of electronic properties are illustrated by reference to a range of materials ...
  29. [29]
    [PDF] Sir Chandrasekhara V. Raman - Nobel Lecture
    It was experimentally discovered that associated with the Rayleigh-Einstein type of molecular scattering, was another and still. Page 4. 270. 1930 C.V.RAMAN.
  30. [30]
    C.V. Raman The Raman Effect - Landmark
    Sir CV Raman discovered in 1928 that when a beam of coloured light entered a liquid, a fraction of the light scattered by that liquid was of a different color.
  31. [31]
  32. [32]
    None
    ### Summary of Raman Selection Rules and Comparison to IR Spectroscopy
  33. [33]
    Surface-enhanced Raman spectroscopy: a half-century historical ...
    Dec 23, 2024 · Abstract. Surface-enhanced Raman spectroscopy (SERS) has evolved significantly over fifty years into a powerful analytical technique.
  34. [34]
    [PDF] Introduction to Nuclear Engineering
    This revision is derived from personal experiences in teaching introductory and advanced level nuclear engineering courses at the undergraduate level.
  35. [35]
    [PDF] DOE Fundamentals Handbook Nuclear Physics and Reactor Theory ...
    In the inelastic scatter reaction, the same neutron/nucleus collision occurs as in elastic scatter. However, in this reaction, the nucleus receives some ...
  36. [36]
    [PDF] Bertram N. Brockhouse - Nobel Lecture
    “Triple-Axis” neutron crystal spectrometer which will be discussed here later. ... as on the triple-axis apparatus, some 210 phonons distributed over two sym-.
  37. [37]
    [PDF] Vibrational-Rotational Spectroscopy
    Typical rotational energies for a diatomic molecule are of the order of 1 to 10 cm-1, while the separation of the vibrational energy levels are typically 1000 ...
  38. [38]
    Vibrational Energy Transfer | The Journal of Physical Chemistry
    Many aspects of the collision dynamics of vibrational energy transfer are presented. Special emphasis is placed on three broad areas within this field.
  39. [39]
    Breakdown of energy transfer gap laws revealed by full-dimensional ...
    Oct 11, 2019 · Numerous experimental and theoretical studies to date have shown that molecular collisions conform to these propensity rules. As a result, most ...
  40. [40]
    Vibrational energy transfer in molecular collisions - AIP Publishing
    State-to-state rate constants for vibrational energy transfer (both V- V and V- n are characterized using an "exponential gap" representation.
  41. [41]
    [PDF] Effusion and Collisions - 5.62 Physical Chemistry II
    Goal: To calculate collision frequency between pairs of molecules in a gas. ... Z = average number of collisions of a single particle per unit time.
  42. [42]
    Relaxation in collisions of vibrationally excited molecules with ...
    Implications for atmospheric chemistry. Chemical ... Rotational energy transfer in H+H2( v ) inelastic collisions ... Relaxation and chemical reaction of ...
  43. [43]
    Close-coupling study of rotational energy transfer of - AIP Publishing
    Quantum close-coupling scattering calculations of rotational energy transfer in the vibrationally excited CO due to collisions with He atom are presented ...
  44. [44]
    The Close-Coupled (CC) equations
    The CC equations arise whenever one describes the scattering of systems with internal degrees of freedom. These equations were introduced into the field of ...
  45. [45]
    Dynamics studied by Quasielastic Neutron Scattering (QENS)
    Feb 2, 2021 · Quasielastic neutron scattering (QENS) allows measurement of the molecular displacements in time and space, from pico- to tens of nanoseconds and from Å ...<|separator|>
  46. [46]
    Insights from quasi-elastic neutron scattering spectroscopy | Journal ...
    Apr 17, 2020 · This tutorial focuses on current and future applications of neutron spectroscopy methods in selectively probing polymer dynamics in complex polymeric systems.
  47. [47]
    Rainbow Scattering in Inelastic Molecular Collisions - SpringerLink
    The role of rainbow scattering in elastic collisions of atoms and atomic ions is well known1,2 and provides an important link between experimental ...
  48. [48]
    The H+H2 reactive system. Progress in the study of the dynamics of ...
    Molecular beam scattering measurements of differential cross sections for D+H2(v=0)→HD+H at Ec.m.=1.5 eV. Source: The Journal of Chemical Physics.
  49. [49]
    Absolute Integral and Differential Cross Sections for the Reactive ...
    To determine reliable threshold energies, it is necessary to account for the thermal motion of the target gas as well as the finite energy distribution of the ...Missing: Ų | Show results with:Ų
  50. [50]
    The effect of van der Waals resonances on reactive cross sections ...
    We found that the van der Waals resonances still survive in integral cross sections even after total angular momentum averaging.
  51. [51]
    [PDF] The Discovery of Quarks* - SLAC National Accelerator Laboratory
    Theoretical analyses of deep inelastic electron-proton scattering made that year by Bjorken (10) suggested that this process might indicate whether there were ...
  52. [52]
    The Discovery of Quarks - Science
    The key evidence for their existence came from a series of inelastic electron-nucleon scattering experiments conducted between 1967 and 1973 at the Stanford ...Missing: original | Show results with:original
  53. [53]
    MINERvA Publications - Fermilab
    “Measurement of inclusive charged-current muon neutrino scattering on hydrocarbon at ⟨Eν⟩∼6 GeV with low three-momentum transfer” Phys.Rev.D 106 (2022) 3, ...
  54. [54]
    Angular Momentum Resolved Inelastic Electron Scattering for ...
    Feb 4, 2025 · Giant resonances (GRs) have emerged as an indispensable source of insights, addressing key questions in nuclear physics and nuclear astrophysics
  55. [55]
    [PDF] Electroexcitation of Giant Resonances between 6.1 MeV and 38 ...
    Jan 1, 2023 · 2 . Inelastic Electron Scattering. Inelastic electron scattering is one of the most powerful tools available for nuclear structure studies ...
  56. [56]
  57. [57]
    Electron-Ion Collider (EIC) - Brookhaven National Laboratory
    The electron beam will reveal the arrangement of the quarks and gluons that make up the protons and neutrons of nuclei. The force that holds quarks together, ...The Machine · Take the INTERACTIVE TOUR · Overview · ePIC CollaborationMissing: November 2025
  58. [58]
    [PDF] Progress and promise of the Electron Ion Collider - CERN Indico
    Aug 26, 2025 · RHIC will conclude operations in 2025. Electron Ion Collider installation will begin after RHIC ops concludes. August 26, 2025. EIC at Lepton- ...
  59. [59]
    Inelastic neutron scattering and lattice-dynamical calculations of
    A strong electron-phonon coupling of in-plane Fe breathing modes has been suggested for the LaO 1 − x F x FeAs superconductor. 17. Meanwhile, a magnetic ...
  60. [60]
    Paramagnons and high-temperature superconductivity in a model ...
    Jun 7, 2022 · Magnetic excitations and phonons simultaneously studied by resonant inelastic x-ray scattering in optimally doped Bi1.5Pb0.55Sr1.6La0.4CuO6+δ.
  61. [61]
    Band gap maps beyond the delocalization limit: correlation between ...
    Jan 16, 2018 · Recent progresses in nanoscale semiconductor technology have heightened the need for measurements of band gaps with high spatial resolution.Results · Reconstructed E Map With... · Methods
  62. [62]
    Using electron energy-loss spectroscopy to measure nanoscale ...
    Aug 1, 2023 · EELS is more sensitive to light elements than photons because the scattering cross-section of electrons is generally larger than that of photons ...
  63. [63]
    From early to present and future achievements of EELS in the TEM
    Jun 24, 2022 · This paper reviews the implementation of Electron Energy Loss Spectroscopy (EELS) in a Transmission Electron Microscope (TEM), as an essential tool for ...
  64. [64]
    Inelastic Neutron Scattering on Polymer Electrolytes for Lithium-Ion ...
    May 7, 2012 · The relationship between ion transport and polymer dynamics is central to the pursuit of solid polymer electrolytes for lithium batteries.Missing: hydrogenous | Show results with:hydrogenous
  65. [65]
    [PDF] Nanosecond Solvation Dynamics in a Polymer Electrolyte for
    The average timescale for solvation dynamics in the polymer electrolyte is one nanosecond, with the motion of polymer segments also around 1 ns.
  66. [66]
  67. [67]
    Using magnetic dynamics to measure the spin gap in a candidate ...
    Feb 1, 2025 · Our observations unveil remarkably slow spin dynamics at the magnetic peak, whose recovery timescale is several nanoseconds. This timescale ...
  68. [68]
    15 years of spin-echo spectroscopy at SNS-NSE: Looking back and ...
    Aug 15, 2025 · This paper reviews the highlights and peculiarities of the SNS-NSE, based at the pulsed neutron source SNS, during its first 1.5 decades of operation.
  69. [69]
    [PDF] 2024 Neutron Scattering Principal Investigator's Meeting
    Keywords: Neutron Spin Echo, Polarized neutron, Quantum Spin Liquids, Topological Insulators. The primary objective of this proposal is to identify, measure ...