Fact-checked by Grok 2 weeks ago

Møller scattering

Møller scattering is the of two identical , e^- + e^- \to e^- + e^-, a fundamental pointlike process in () mediated by the exchange of virtual photons. Named after Danish Christian Møller, who derived the relativistic cross-section in his 1931 paper considering retardation effects in particle collisions, it represents an early application of to electron interactions. The process is indistinguishable for the two incoming electrons, requiring antisymmetrized wave functions and leading to two contributing Feynman diagrams at lowest order: the direct (t-channel) and exchange (u-channel) photon exchanges. This second-order QED calculation yields a differential cross-section that accounts for spin-averaged amplitudes and has been extended to higher orders, including next-to-leading and next-to-next-to-leading corrections for precision comparisons. Gauge invariance ensures the amplitude's consistency across different gauges, highlighting QED's robustness. Historically, Møller's work addressed challenges in understanding high-energy electron passage through matter, predating full QED renormalization and providing experimental tests in the 1930s through cloud chamber observations. Despite initial semi-classical approaches, it transitioned into perturbative QED frameworks, influencing developments in particle scattering theory. In modern physics, Møller scattering is crucial for validating QED at low energies, as demonstrated by precise measurements at 2.5 MeV using accelerators, which agree with theoretical predictions including radiative corrections. It also probes electroweak interactions via parity-violating asymmetries, enabling determinations of the weak mixing angle in experiments like the planned MOLLER at Jefferson Lab, where electron beam polarization enhances sensitivity to beyond-Standard-Model effects. These applications underscore its role in testing the Standard Model and constraining new physics.

Introduction

Definition and Physical Process

Møller scattering is the elastic scattering process of two identical electrons, denoted as e^- e^- \to e^- e^-, first theoretically described in the context of for the passage of high-energy electrons through matter. This interaction represents a fundamental pointlike process in , where the electrons do not change their identity or produce additional particles, preserving the total number of electrons before and after the collision. In the physical process, the two incoming electrons interact via the exchange of a , which mediates the electromagnetic force and transfers between them without real emission or absorption at the lowest order. This exchange leads to a deflection of the electron trajectories, with the virtual photon's space-like ensuring the process remains and compliant with 's gauge invariance. The indistinguishability of the electrons as fermions necessitates an antisymmetric of the interaction, incorporating both direct and exchange contributions to the to satisfy the . Unlike , which involves an electron and a positron (e^- e^+ \to e^- e^+) and includes both photon exchange and channels due to the distinguishable particles, Møller scattering exclusively features photon exchange between identical fermions, requiring antisymmetrization and excluding annihilation processes. This distinction arises from the fermionic nature of electrons, leading to effects not present in the boson-mediated or distinguishable particle cases. Basic kinematics of Møller scattering involve two incoming with relativistic energies E_1 and E_2, typically considered in the laboratory frame where one electron may be at rest or both in a setup, scattering into final states at angles \theta relative to the initial directions. The identical particle implies that scattering angles \theta and $180^\circ - \theta are indistinguishable, affecting the and emphasizing the need for symmetric kinematic descriptions in experimental analyses.

Historical Development and Importance

Møller scattering was first theoretically described by Danish physicist Christian Møller in 1932 as part of his investigation into the passage of high-speed electrons through matter, where he derived the relativistic quantum mechanical formula for electron-electron collisions using Dirac's hole theory framework. This work addressed the scattering process in the context of early , providing a foundational calculation for interactions between identical electrons and highlighting the role of exchange effects due to their fermionic nature. The theory evolved significantly in the 1940s with the renormalization and full formulation of QED by Richard Feynman, Julian Schwinger, and Sin-Itiro Tomonaga, who incorporated Møller's scattering into the perturbative framework, treating it as the first non-trivial QED process involving two electrons beyond simpler cases like Compton scattering. Feynman's space-time approach, in particular, used Møller scattering as a prototype to develop diagrammatic rules, ensuring consistency with gauge invariance and relativistic invariance in higher-order corrections. Møller scattering holds central importance as a precision test of at low energies, where experimental measurements confirm theoretical predictions to high accuracy, validating principles such as identical particle statistics through antisymmetrization of amplitudes and the maintenance of gauge invariance in interactions. In modern contexts, it forms the basis for modeling multi- interactions in relativistic plasmas, such as those in laser-plasma acceleration schemes, and in particle accelerators, where it governs intra-beam in storage rings, affecting and beam stability. Additionally, polarized Møller scattering enables precision measurements of parity-violating effects, as in the MOLLER experiment at Jefferson Lab, which began construction in early 2025 and aims to determine the weak mixing angle with unprecedented accuracy, with first physics results anticipated by mid-2027.

Theoretical Foundations

Kinematics of Electron-Electron Scattering

In electron-electron scattering, known as Møller scattering, the kinematics are described using Lorentz-invariant , which capture the energy and momentum transfers in a frame-independent manner. The variable s = (p_1 + p_2)^2 represents the square of the total center-of-mass energy available for the process, where p_1 and p_2 are the of the incoming s. The variable t = (p_1 - p_3)^2 denotes the square of the transfer in the direct scattering channel, with p_3 being the of one outgoing electron, and is typically negative for physical . Similarly, u = (p_1 - p_4)^2 corresponds to the momentum transfer in the exchange channel, where p_4 is the of the other outgoing electron. These variables satisfy the relation s + t + u = 4m^2, where m is the , arising from four-momentum conservation and the on-shell conditions for massive particles. The center-of-mass (CM) frame provides a convenient reference for analyzing the scattering geometry, where the total three-momentum vanishes: \vec{p}_1 + \vec{p}_2 = 0 = \vec{p}_3 + \vec{p}_4. In this frame, the incoming electrons have equal and opposite momenta, and the scattering is characterized by the polar \theta between an incoming electron's momentum and one of the outgoing electrons' momenta. The magnitudes of the incoming and outgoing momenta are related through s, with |\vec{p}| = \sqrt{ [s - 4m^2](s)/ (4s) } for elastic scattering, ensuring . This frame simplifies the description of angular distributions, as the differential cross-section depends primarily on \theta and the relativistic invariants. In contrast, the (lab) frame often features unequal incoming energies, such as when one is at rest or has negligible relative to the other, complicating direct calculations. Transformations between the lab and frames are achieved via Lorentz boosts along the beam direction, preserving the while adjusting angles and energies; for instance, the scattering angle \theta relates to the lab angle through the boost velocity derived from s and the lab energies. Such transformations are essential for interpreting experimental data, where measurements are typically performed in the lab frame. Relativistic effects dominate in Møller scattering at high energies (E \gg m), where the formulation remains Lorentz-invariant through the , allowing approximations that neglect the in kinematic relations. In the massless limit, the relation simplifies to s + t + u = 0, and momentum transfers are expressed as t = - (s/2) (1 - \cos \theta) in the frame, highlighting the forward-peaking nature of the process due to exchange. These approximations are valid for energies far exceeding the rest mass, as encountered in particle accelerators.

Role in Quantum Electrodynamics

Møller scattering serves as a cornerstone process in , illustrating the interaction between two electrons through the exchange of a . In QED, the behavior of electrons is governed by the , which describes their relativistic quantum dynamics as fermions, while the photon mediation follows from the quantized Maxwell equations incorporated into the . This framework captures the electromagnetic force as arising from the coupling between the electron's Dirac field and the field via the minimal substitution in the . As electrons are identical fermions, the mandates that the total be antisymmetric under particle exchange. Consequently, the for Møller scattering must incorporate antisymmetrization, resulting in destructive interference between the direct photon exchange (t-channel diagram) and the exchange process (u-channel diagram) where the outgoing electrons are interchanged. This quantum interference effect distinguishes Møller scattering from scattering involving distinguishable particles, such as electron-muon scattering, and highlights the role of identical particle statistics in predictions. The process upholds the U(1) gauge invariance inherent to , ensuring that physical observables remain unchanged under gauge transformations of the . Ward identities, derived from this symmetry, enforce relations among vertex functions and propagators that guarantee the transversality of the and the consistency of amplitudes, preventing unphysical longitudinal contributions. These identities are crucial for the renormalizability and predictive power of in calculations involving Møller scattering. Møller scattering represents the quantum relativistic extension of classical , which describes non-relativistic Coulomb repulsion between charged particles. Christian Møller's 1932 calculation generalized the Rutherford formula to include relativistic kinematics and quantum effects, such as electron spin interactions and exchange contributions, while recovering the classical limit in the low-energy, non-relativistic regime without spin polarization. This bridge underscores how resolves classical divergences and incorporates finite-size effects absent in Rutherford's point-particle model.

Feynman Diagrams

Direct and Exchange Processes

In , Møller scattering at tree level is represented by two primary Feynman diagrams that capture the between identical electrons: the and the . These diagrams illustrate the exchange of a between the incoming electrons, with the corresponding to the t-channel and the to the u-channel, where t and u are representing momentum transfers. The antisymmetrized total amplitude arises from their coherent sum, essential for fermions to satisfy the . The direct process depicts a t-channel in which one incoming , represented by a Dirac spinor u(p_1), emits a at a coupled via -i e \gamma^\mu, which propagates with denominator q^2 (where q = p_1 - p_3 is the transfer) before being absorbed by the second incoming , also a Dirac spinor u(p_2), at another with the same coupling. This structure embodies a classical-like electrostatic repulsion between the charged particles, favoring small-angle forward where the transfer |t| is minimal and the term diverges softly. In Møller's original relativistic treatment, this direct corresponds to the unaltered trajectories without particle swap. The process, conversely, is a u-channel that enforces the indistinguishability of electrons by interchanging their outgoing states: the first electron's u(p_1) now connects to the leading to outgoing u(p_4), with the $1/q^2 (now q = p_1 - p_4) linking to the second electron's lines u(p_2) to u(p_3), again via -i e \gamma^\mu . This configuration introduces a quantum mechanical term where the electrons effectively swap identities, crucial for the overall antisymmetric under . Physically, it generates with the direct process, producing diffraction-like patterns in the distribution and suppressing close encounters due to quantum statistics, beyond classical expectations. Møller's formulation highlighted this as a distinct to account for identical particle effects in relativistic collisions.

Invariant Amplitudes

In Møller scattering, the Lorentz-invariant matrix elements are obtained from the t-channel direct and the u-channel exchange , accounting for the indistinguishability of the identical . The matrix element for the direct process is \mathcal{M}_\text{direct} = -\frac{e^2}{t} \bar{u}(p_3) \gamma^\mu u(p_1) \bar{u}(p_4) \gamma_\mu u(p_2), where p_1, p_2 are the incoming electron four-momenta, p_3, p_4 are the outgoing ones, t = (p_1 - p_3)^2, e is the , and the Dirac spinors u(p) satisfy the normalization \bar{u}(p) u(p) = 2m for massive electrons with mass m. Similarly, the exchange matrix element is \mathcal{M}_\text{exchange} = -\frac{e^2}{u} \bar{u}(p_4) \gamma^\mu u(p_1) \bar{u}(p_3) \gamma_\mu u(p_2), with u = (p_1 - p_4)^2. The total antisymmetrized amplitude is \mathcal{M} = \mathcal{M}_\text{direct} - \mathcal{M}_\text{exchange}, where the minus sign arises from Fermi-Dirac statistics. The spin-averaged squared amplitude, required for unpolarized cross-section calculations, is computed as |\mathcal{M}|^2 = \frac{1}{4} \sum_\text{spins} |\mathcal{M}|^2, averaging over the two initial electron spins and summing over the final ones. This involves expanding the square using the completeness relation \sum_s u(p) \bar{u}(p) = \slash{p} + m and evaluating traces of Dirac matrices. In the high-energy (massless) limit, the result is |\mathcal{M}|^2 = 2 e^4 \left[ \frac{s^2 + u^2}{t^2} + \frac{s^2 + t^2}{u^2} - \frac{2 s^2}{t u} \right], where s = (p_1 + p_2)^2 is the Mandelstam center-of-mass energy squared, and the interference term originates from the cross product \mathcal{M}_\text{direct} \mathcal{M}_\text{exchange}^*. The factors of $2E in the relativistic normalization of the states ensure that |\mathcal{M}|^2 is Lorentz invariant under boosts.

Cross-Section Computation

Differential Cross-Section Formula

The differential cross-section for Møller scattering, which describes the angular distribution of elastically scattered electrons in (QED), is derived from the spin-averaged squared matrix element |\mathcal{M}|^2 obtained from the direct and exchange Feynman diagrams, combined with appropriate kinematic factors in the center-of-mass (CM) frame. For unpolarized electrons, the general expression in terms of s, t, and u (where s is the CM energy squared, t is the momentum transfer squared in the t-channel, and u in the u-channel) is \frac{d\sigma}{d\Omega} = \frac{\alpha^2}{4s} \left[ \frac{s^2 + u^2}{t^2} + \frac{s^2 + t^2}{u^2} - \frac{4s^2}{tu} \right], where \alpha = e^2/(4\pi) is the fine-structure constant in natural units (\hbar = c = 1), and the negative sign in the interference term reflects the antisymmetric wave function for identical fermions (noting that t < 0 and u < 0). This formula incorporates the flux factor and phase space for two-body scattering, with the matrix element averaged over initial spins (factor of $1/4) and summed over final spins, following the standard QED prescription for $2 \to 2 processes. In the high-energy limit where the electron mass m_e \to 0 (valid for CM energies much greater than m_e c^2 \approx 0.511 MeV), the kinematics simplify with s \approx 4E^2 (where E is the CM energy per electron), t \approx -s \sin^2(\theta/2), and u \approx -s \cos^2(\theta/2), with \theta the CM scattering angle. More commonly, this is expressed directly in angular variables as \frac{d\sigma}{d\Omega} = \frac{\alpha^2}{s} \left[ \frac{1 + \cos^4(\theta/2)}{\sin^4(\theta/2)} + \frac{1 + \sin^4(\theta/2)}{\cos^4(\theta/2)} - \frac{2(1 + \cos^2(\theta/2) \sin^2(\theta/2))}{\sin^2(\theta/2) \cos^2(\theta/2)} \right], confirming the core QED prediction with destructive from the . The angular distribution exhibits strong forward (\theta \to 0) and backward (\theta \to 180^\circ) peaking due to the $1/t^2 and $1/u^2 poles from the photon exchange in the t- and u-channels, respectively, with the interference term suppressing scattering at intermediate angles and ensuring under \theta \to 180^\circ - \theta for particles. For finite masses, the full relativistic expression includes additional terms proportional to m_e^2/s, but these are negligible at high energies and restore unitarity away from the poles. This differential cross-section serves as the fundamental observable for testing vertex structure and radiative corrections in experiments.

Integration to Total Cross-Section

The total cross-section for Møller scattering is obtained by integrating the cross-section over all angles in the center-of-mass : \sigma_\text{total} = \int \frac{d\sigma}{d\Omega}\, d\Omega = 2\pi \int_0^\pi \frac{d\sigma}{d\Omega} \sin\theta \, d\theta. This diverges in the regime due to the forward as t \to 0 (corresponding to \theta \to 0), where the leads to a $1/t^2 term in the squared matrix element, producing a $1/\theta^4 behavior in d\sigma/d\Omega. Regularization of this is achieved by imposing a minimum scattering angle cutoff \theta_\text{min} to exclude unphysical forward , or through soft resummation, where infrared divergences from virtual loop corrections cancel against those from real soft bremsstrahlung emission below an \Delta E. The resummation introduces a soft involving logarithms of \Delta E / \mu (with \mu as a fictitious mass regulator), yielding a finite inclusive cross-section. In the high-energy limit (s \gg m_e^2), the regularized total cross-section takes the approximate form \sigma \approx \frac{8\pi \alpha^2}{3 s} \ln\left( \frac{s}{m_e^2} \right), where the logarithmic enhancement reflects the scale separation between the center-of-mass energy and the electron mass acting as a natural infrared regulator for collinear emissions. In the non-relativistic low-energy limit, the total cross-section approximates \sigma \approx \frac{8\pi \alpha^2}{3 m_e v^4}, connecting the quantum electrodynamic result to the classical for identical charged particles with m_e/2 and relative velocity v. Numerical evaluation of the presents challenges due to the identical fermions in the final state; double-counting is avoided by restricting the angular integration to $0 \leq \theta \leq \pi/2 (exploiting the d\sigma/d\Omega (\theta) = d\sigma/d\Omega (\pi - \theta)) and multiplying by a factor of 2, while the differential cross-section itself already antisymmetrizes the direct and exchange contributions.

Applications and Experimental Aspects

Use in Particle Physics Experiments

Møller scattering has been experimentally investigated in fixed-target setups using high-energy electron beams incident on unpolarized or polarized targets, providing a clean probe of - interactions. A seminal experiment was the E158 parity-violation study at the Stanford Linear Accelerator Center (SLAC), where a 45-50 GeV polarized beam was scattered off a target, achieving high precision in measuring the weak mixing angle through the parity-violating asymmetry in the scattering cross-section. This setup utilized a spectrometer to detect scattered electrons at forward angles, demonstrating the feasibility of Møller scattering for electroweak precision tests at GeV scales. Similar configurations have been employed in precursor experiments to larger colliders, adapting Møller processes for beam diagnostics in linear accelerators and storage rings. In experiments, Møller scattering serves as a reliable tool for determining , leveraging the spin-dependent in the differential cross-section. Møller polarimeters typically involve the electrons from polarized electrons in a thin target, such as magnetized iron foil or gas, where the analyzing power approaches unity at specific , enabling absolute measurements with uncertainties below 1%. For instance, dedicated Mott polarimeters have been integrated with Møller setups to study transfer in relativistic , as demonstrated in experiments at facilities like the Microtron, achieving high accuracy for transverse polarizations up to 85%. This application is crucial for polarized experiments, ensuring reliable control in colliders and fixed-target setups. Møller scattering also finds application in luminosity monitoring within lepton collider environments, where it complements small-angle Bhabha scattering by providing an independent measure of beam intensity through correlated event rates in detector systems. In the SLAC E158 apparatus, Møller events were cross-checked against dedicated luminosity monitors to account for beam fluctuations and density effects, validating the integrated luminosity to better than 1% precision. Furthermore, it enables rigorous tests of quantum electrodynamics (QED) at high energies, with measurements up to TeV scales probing radiative corrections and electroweak interference; for example, electroweak corrections in Møller processes at √s ≈ 2 TeV can contribute up to 30% to the asymmetry, offering sensitivity to new physics beyond the Standard Model. Recent developments emphasize polarized Møller scattering for studying parity violation, particularly in the MOLLER experiment at Jefferson Laboratory, which uses an 11 GeV polarized beam on a liquid hydrogen target to measure the parity-violating asymmetry with unprecedented 0.7% precision, aiming to refine the weak mixing angle by a factor of five over prior results. As of 2025, the MOLLER project is in the construction phase, with critical design reviews completed, assembly beginning in early 2025, commissioning targeted for late 2026, and first physics data anticipated in early 2027, incorporating advanced quartz detectors and a spectrometer to handle high event rates while minimizing backgrounds from beamline scattering. This experiment builds on the SLAC E158 legacy, extending low-energy electroweak tests to constrain TeV-scale physics models.

Comparisons with Observations

Early experimental verifications of Møller scattering predictions were conducted using beta particles from radioactive sources. In 1932, F. C. Champion measured the energy distribution of scattered electrons from radium E (²¹⁰Bi) beta decay, demonstrating agreement between the observed scattering and the relativistic Møller formula derived from Dirac theory, within the experimental resolution of the time. This work provided one of the first confirmations of electron-electron scattering in the non-relativistic to mildly relativistic regime. Accelerator-based experiments in the mid-20th century extended these tests to higher energies. A landmark measurement at the Stanford Linear Accelerator in 1966 examined the differential cross-section for electron-electron scattering at a center-of-mass energy of 1.112 GeV. The results agreed with tree-level predictions to within 5%, accounting for experimental uncertainties, and helped validate the theory in the relativistic domain where higher-order corrections become relevant. Modern low-energy experiments continue to probe QED validity with improved precision. In a 2020 measurement at 2.5 MeV incident energy, momentum spectra of scattered electrons were recorded at laboratory angles of 30°, 35°, and 40° using thin foil targets and detectors. The observed spectra matched radiative simulations (including effects) to within 1.5% across all angles, with statistical and systematic uncertainties at the percent level, confirming the theory's accuracy even where the electron mass is comparable to the energy scale. High-energy parity-violating Møller scattering experiments further test as part of electroweak validations, with no significant deviations observed. The SLAC E158 experiment in 2005 measured the parity-violating asymmetry at 45–50 GeV beam energy, yielding a value of -131 ± 16 ppm consistent with the prediction of -114 ppm to 1.5σ, relying on QED cross-sections accurate to better than 0.5% after radiative corrections. Radiative corrections up to order α³, including virtual and soft effects, have been incorporated in these comparisons and match experimental data without anomalies; for instance, hard contributions alter the cross-section by up to 10% at forward angles but are well-reproduced in simulations. These agreements underscore QED's robustness across energy scales from MeV to GeV.

References

  1. [1]
    [1903.09265] Measurement of Moller Scattering at 2.5 MeV - arXiv
    Mar 21, 2019 · Moller scattering is one of the most fundamental processes in QED. Understanding it to high precision is necessary for a variety of modern ...
  2. [2]
    Møller scattering: a neglected application of early quantum ...
    Roqué, X. Møller scattering: a neglected application of early quantum electrodynamics. Arch. Hist. Exact Sci. 44, 197–264 (1992).
  3. [3]
    [PDF] Chapter 5 Elements of quantum electrodynamics - ITP Lecture Archive
    Example We treat the case of Møller scattering e−e− → e−e− as a typical example for the application of the Feynman rules. We first define our initial and final ...
  4. [4]
    [2107.12311] Møller scattering at NNLO - arXiv
    Jul 26, 2021 · We present a calculation of the full set of next-to-next-to-leading-order QED corrections to unpolarised Møller scattering.
  5. [5]
    Parity violation in Møller scattering
    Jul 23, 2025 · We include electroweak effects in Moller scattering at low energies in an effective field theory approach and compute the left-right parity- ...
  6. [6]
  7. [7]
    [PDF] arXiv:1903.09265v2 [nucl-ex] 13 Apr 2019
    Apr 13, 2019 · Møller (electron-electron) scattering is a fundamental, purely-pointlike process in QED, which provides an im- portant means to test the ...
  8. [8]
    [PDF] 129A Lecture Notes - Quantum ElectroDynamics 1 Quantum ...
    If two electrons scatter, called Møller scattering, they exchange a virtual photon. But two electrons are identical, and you can draw another diagram where ...
  9. [9]
    Zur Theorie des Durchgangs schneller Elektronen durch Materie
    Zur Theorie des Durchgangs schneller Elektronen durch Materie. Chr. Møller,. Chr. Møller. Kopenhagen, Universitetets Institut for teoretisk Fysik. Search for ...
  10. [10]
    Measurement of Møller scattering at 2.5 MeV
    Jul 16, 2020 · Møller (electron-electron) scattering is a fundamental, purely pointlike process in QED, which provides an important means to test the Standard ...
  11. [11]
    Storage ring lattice calibration using resonant spin depolarization
    Jul 1, 2013 · Møller scattering polarimetry. Møller scattering is electron-electron scattering and occurs within a bunch in the storage ring. The ...
  12. [12]
    [PDF] Chapter 2 Relativistic kinematics
    Figure 2.6: t-channel. It is called “s-channel” reaction, because the only positive Mandelstam variable is s. Ts describes the scattering dynamics of the ...
  13. [13]
    [PDF] Particle Physics 2011
    Oct 13, 2011 · In QFT interaction of 2 electrons by exchange of virtual photon is described by a perturbative series hoton is described by a perturbative ...
  14. [14]
    [PDF] PARTICLE PHYSICS
    Møller scattering. ... Comment physically the resulting formula. 3. General kinematics relation.- Demonstrate that the sum of the three Mandelstam variables is.<|control11|><|separator|>
  15. [15]
    [PDF] The basics of Quantum Electrodynamics, and its role in Particle ...
    The reaction e−e− → e−e− is called Møller scattering [4]. There are two Feynman diagrams for Møller scattering, shown in Figure 2.
  16. [16]
    Electron-Electron Scattering
    The relative minus sign between the direct and exchange terms is due to the Fermi statistics, which requires the amplitude to be antisymmetric under interchange ...
  17. [17]
    [PDF] arXiv:1411.4088v2 [nucl-ex] 3 Dec 2014
    Dec 3, 2014 · MOLLER proposes to measure the parity-violating asymmetry in the scattering of longitudinally polarized electrons off unpolarized electrons, ...
  18. [18]
    [PDF] Ward Identities - UT Physics
    First of all, the Ward identity (5) provides for the gauge invariance of the scattering amplitudes (2) and (4) despite the gauge-dependence of the photonic ...Missing: Møller | Show results with:Møller
  19. [19]
    Electron Scattering Processes in Non-Monochromatic and ... - MDPI
    Jul 1, 2004 · The second problem was analyzed by the Danish physicist Christian Møller [15,16] during 1931–1932, who also derived a scattering formula for the ...
  20. [20]
    [PDF] The MOLLER Experiment - Experimental Hall A
    Sep 12, 2011 · The leading order Feynman diagrams relevant for Møller scattering, involving both direct and exchange diagrams that interfere with each ...
  21. [21]
    [PDF] Dylan J. Temples: Moller Scattering - New IAC Wiki
    Feb 16, 2017 · We recover, in the ultra-relativistic limit (i.e., massless electron):. | ¯Mt|2 = 2e4 s2 + u2 t2 . (36). Page 3 of 12. Page 4. Dylan J. Temples.<|control11|><|separator|>
  22. [22]
    [PDF] Chapter 5
    ... Møller scattering formula, one can readily obtain the e-e+ → e- e+ Bhabha scattering formula by crossing (s ↔ u). The scattering amplitudes for e- e- → e ...
  23. [23]
    Quadratic electroweak corrections for polarized Møller scattering
    Jan 10, 2012 · BREMSSTRAHLUNG AND CANCELLATION OF INFRARED DIVERGENCES. To evaluate the cross section induced by the emission of one soft photon with energy ...
  24. [24]
    [PDF] Radiative Corrections to Fixed Target Møller Scattering Including ...
    We find that the majority of the QED corrections to ALR cancel when both virtual and hard photon corrections are added, as anticipated in [10, 11]; they ...
  25. [25]
    [PDF] arXiv:1110.1750v1 [hep-ph] 8 Oct 2011
    Oct 8, 2011 · Polarized Møller scattering has been a well-studied low-energy reaction for close to eight decades now [1], but has attracted especially active ...
  26. [26]
    Nonrelativistic electron–electron Møller scattering in a nonadiabatic ...
    Jan 29, 2020 · Notably, the minimum cross-section occurs at θs = 26°, indicating that the Møller nonrelativistic scattering of SE from the PE alone is not ...
  27. [27]
    [PDF] A Precision Measurement of the Weak Mixing Angle in Møller ...
    Calculations indicate that the pion background passing through the. Møller detector is likely to have an asymmetry at the level of 1 ppm, an order of magnitude.Missing: original title
  28. [28]
    Precision Measurement of the Weak Mixing Angle in Møller Scattering
    The principal components of the experimental apparatus are the polarized electron beam, beam diagnostics, the liquid hydrogen target, the spectrometer/ ...
  29. [29]
    Accurate Determination of the Electron Spin Polarization In ... - arXiv
    Mar 21, 2022 · Møller polarimetry uses the double polarized scattering asymmetry of a polarized electron beam on a target with polarized atomic electrons.
  30. [30]
    Measurement of polarization transfer in M\o{}ller scattering of ...
    Nov 29, 2021 · A dedicated Mott polarimeter was designed to study the final spin state in Møller scattering of transversely polarized electron beams in the relativistic ...
  31. [31]
    [PDF] First Observation of the Parity Violating Asymmetry in Moller Scattering
    Density fluctuations due to electron beam were studied by looking at the residual correlation between the. Moller detector and luminosity monitor (described ...
  32. [32]
    [PDF] NLO and NNLO EWC for PV Møller Scattering - CERN Indico
    For the high-energy region √s ∼ 2000 GeV the contribution of the quadratic EWC can reach +30%. For the two-loop relative corrections, we can say that the ...
  33. [33]
    The MOLLER Experiment: An Ultra-Precise Measurement of ... - arXiv
    Nov 15, 2014 · The proposed MOLLER experiment will improve on this result by a factor of five, yielding the most precise measurement of the weak mixing angle at low or high ...
  34. [34]
    [PDF] An Overview of the MOLLER Experiment at Jefferson Lab - Indico
    Mar 17, 2025 · Summary. • MOLLER provides a high-precision test of electroweak theory in a low-energy setting. -Can help resolve discrepancies in sin! 𝜃".
  35. [35]
    (PDF) The MOLLER Experiment: An Ultra-Precise Measurement of ...
    Aug 1, 2025 · The MOLLER Experiment: An Ultra-Precise Measurement of the Weak Mixing Angle Using Møller Scattering. Zuhal Seyma Demiroglu. 1. Introduction.
  36. [36]
    The scattering of fast β-particles by electrons
    The present paper gives an account of an investigation of the scattering of fast (3-particles by electrons, using the expansion method. Among the many formulae!Missing: original | Show results with:original
  37. [37]
    Test of Quantum Electrodynamics by Electron-Electron Scattering
    Phys. Rev. Lett. 16, 1127 (1966) ... Test of Quantum Electrodynamics by Electron-Electron Scattering. W. C. Barber · B. Gittelman and G. K. O'Neill · B. Richter.
  38. [38]
    Test of Quantum Electrodynamics by Electron-Electron Scattering
    Jun 1, 1971 · We have measured the differential cross section for electron-electron scattering at a total energy of 1112 MeV in the center-of-mass system.