Fact-checked by Grok 2 weeks ago

Exchange force

The exchange force, more precisely termed the , is a quantum mechanical effect that arises from the indistinguishability of identical particles. According to the principles of quantum statistics, the total of a of identical fermions must be antisymmetric under particle , while for bosons it must be symmetric. This requirement leads to correlation effects in the , manifesting as an effective interaction energy—often attractive or repulsive depending on and spatial overlaps—that influences the behavior of multi-particle s. This interaction plays a crucial role in various physical contexts, including the stability of atomic electron shells, covalent bonding in molecules, magnetic ordering in solids, and the binding of nucleons in atomic nuclei. In nuclear physics, it contributes to modeling the strong nuclear force, which has a short range of approximately 1–2 femtometers, through both statistical exchange effects and the exchange of virtual mesons. The concept originated in the late 1920s with Werner Heisenberg's work on quantum mechanics for multi-electron atoms, introducing exchange effects to satisfy wave function symmetry for identical fermions. In 1932, following James Chadwick's discovery of the neutron, Heisenberg extended these ideas to nuclear structure, proposing that the strong force between protons and neutrons arises from their exchange, similar to electron exchange in molecules, incorporating natural saturation. In 1933, Ettore Majorana refined the theory by incorporating spatial exchange, aiding explanations of nuclear saturation and scattering. The 1935 breakthrough by Hideki Yukawa introduced meson exchange via a massive particle (mass ~200 electron masses), yielding the Yukawa potential V(r) = -g^2 \frac{e^{-\mu r}}{r} (with \mu tied to meson mass and g the coupling), matching the short-range attraction; pions were discovered in 1947, confirming this. Characteristics include charge independence (equal action on like and unlike pairs), spin-spatial dependence with tensor components, and a blend of direct and exchange terms (e.g., ~50% in Serber models) to fit data. Modern views frame it as an effective theory linking at scales to nuclear phenomena.

Foundations in

Identical Particles and Symmetry

In , identical particles, also known as , are those of the same species whose individual identities cannot be distinguished by any physical measurement, as they share identical intrinsic properties such as mass and charge. This indistinguishability arises fundamentally from the principles of , where the state of the system is described by a wavefunction that does not permit labeling particles with unique trajectories or identifiers. In contrast, treats particles as distinguishable entities, allowing for unique labeling based on initial conditions or positions, which enables straightforward of their roles without altering the physical description. The quantum mechanical treatment of such particles requires the total wavefunction to exhibit definite under particle , dictated by the symmetrization postulate. For bosons, the wavefunction must be totally symmetric upon interchanging any two particles, while for fermions, it must be totally antisymmetric. This classification is governed by the spin-statistics theorem, which establishes that particles with integer (0, 1, 2, ...) obey Bose-Einstein statistics and thus symmetric wavefunctions, whereas those with , 3/2, ...) follow Fermi-Dirac statistics and antisymmetric wavefunctions. These symmetry requirements have profound physical implications. Bosons, such as photons with , can occupy the same simultaneously, facilitating phenomena like Bose-Einstein condensation where large numbers of particles "bunch" into the . Fermions, like electrons with , cannot share the same state due to the vanishing of the antisymmetric wavefunction when attempting to do so, which enforces spatial separation and underpins the . The P_{12}, which interchanges the coordinates and spins of particles 1 and 2, formalizes this: for a valid two-particle wavefunction \Psi(1,2), it satisfies \Psi(1,2) = \pm P_{12} \Psi(1,2), with the plus sign for bosons and minus for fermions. A illustrative example is two non-interacting identical particles confined to a one-dimensional (a ""). The single-particle eigenfunctions are \psi_n(x) = \sqrt{2/L} \sin(n \pi x / L) for quantum numbers n = 1, 2, \dots and box length L. For distinct states n \neq m, the bosonic two-particle wavefunction is the symmetric combination: \frac{1}{\sqrt{2}} \left[ \psi_n(x_1) \psi_m(x_2) + \psi_m(x_1) \psi_n(x_2) \right], while the fermionic one is antisymmetric: \frac{1}{\sqrt{2}} \left[ \psi_n(x_1) \psi_m(x_2) - \psi_m(x_1) \psi_n(x_2) \right]. If n = m, the fermionic combination vanishes identically, prohibiting both particles from occupying the same state, whereas the bosonic case allows it with the unsymmetrized form \psi_n(x_1) \psi_n(x_2). This demonstrates how enforces distinct behavioral patterns without invoking interactions.

Fermions and the Pauli Exclusion Principle

Fermions are particles characterized by half-integer spin values, such as electrons, protons, and neutrons, which must obey specific symmetry requirements in quantum mechanics. The Pauli exclusion principle, formulated by Wolfgang Pauli in 1925, states that no two identical fermions can occupy the same quantum state simultaneously. This principle emerges directly from the antisymmetry of the total wavefunction for a system of identical fermions, ensuring that exchanging any two particles changes the sign of the wavefunction. For a system of two electrons, dictates the form of the wavefunction based on spin configuration. If the spins form a symmetric (total spin 1), the spatial part of the wavefunction must be antisymmetric to maintain overall antisymmetry. Conversely, for an antisymmetric spin state (total spin 0), the spatial wavefunction is symmetric. This coupling between spin and spatial enforces the exclusion by prohibiting configurations where both electrons share identical spatial and spin quantum numbers. A key consequence of this antisymmetry is the formation of an "exchange hole" in the density distribution. The exchange hole reduces the probability of finding two fermions with parallel spins at the same location, effectively creating a repulsive even in the absence of classical forces. This phenomenon arises purely from the wavefunction's symmetry and underlies many quantum effects in multi-particle systems. In , the governs the filling of shells, allowing at most two electrons per orbital with opposite spins to satisfy antisymmetry. This arrangement explains the periodic table's structure and prevents all electrons from collapsing into the lowest energy state. Furthermore, the principle ensures the stability of matter by counteracting attractive forces; without it, electrons could occupy arbitrarily small volumes, leading to gravitational or electromagnetic collapse, as seen in the bounded energy of atomic systems.

Mathematical Formulation

Wavefunction Antisymmetrization

In , the wavefunctions of identical fermions must be antisymmetric under the exchange of any two particles to satisfy the . This requirement ensures that the total wavefunction vanishes if two fermions occupy the same state. For a system of two identical fermions, the antisymmetric spatial wavefunction is constructed as a of the individual orbital products that changes sign upon particle interchange: \Psi_A(1,2) = \frac{1}{\sqrt{2}} \left[ \phi_a(1)\phi_b(2) - \phi_a(2)\phi_b(1) \right], where \phi_a and \phi_b are single-particle spatial orbitals, and the labels 1 and 2 denote the particles' coordinates. This form guarantees antisymmetry: \Psi_A(2,1) = -\Psi_A(1,2). The normalization of this wavefunction requires accounting for the overlap between the orbitals. Assuming the individual orbitals are normalized (\int |\phi_a|^2 dV = 1, \int |\phi_b|^2 dV = 1), the exact normalization factor is \frac{1}{\sqrt{2(1 - S^2)}}, where S = \int \phi_a^* \phi_b \, dV is the overlap integral. This ensures \int |\Psi_A|^2 dV_1 dV_2 = 1. If the orbitals are orthogonal (S = 0), the simpler factor \frac{1}{\sqrt{2}} suffices. Such antisymmetric wavefunctions are orthogonal to symmetric ones and to those with identical orbitals, as \Psi_A(1,1) = 0. For N identical fermions, the antisymmetric wavefunction is generalized using a : \Psi(1, \dots, N) = \frac{1}{\sqrt{N!}} \det \begin{bmatrix} \phi_1(1) & \phi_2(1) & \cdots & \phi_N(1) \\ \phi_1(2) & \phi_2(2) & \cdots & \phi_N(2) \\ \vdots & \vdots & \ddots & \vdots \\ \phi_1(N) & \phi_2(N) & \cdots & \phi_N(N) \end{bmatrix}, where the \phi_i are orthonormal spin-orbitals (including spatial and spin parts). This determinant automatically enforces antisymmetry under any particle and vanishes if any two spin-orbitals are identical. For the two-particle case, the Slater determinant reduces to the explicit form: \Psi_A(1,2) = \frac{1}{\sqrt{2!}} \det \begin{bmatrix} \phi_a(1) & \phi_b(1) \\ \phi_a(2) & \phi_b(2) \end{bmatrix} = \frac{1}{\sqrt{2}} \left[ \phi_a(1)\phi_b(2) - \phi_a(2)\phi_b(1) \right]. This construction was introduced by John C. Slater in 1929 to represent multi-electron wavefunctions in atomic spectra. The total wavefunction for fermions must also incorporate spin degrees of freedom, typically as a product of spatial and spin parts: \Psi_\text{total} = \Psi_\text{spatial} \times \chi_\text{spin}, ensuring overall antisymmetry. For two spin-1/2 particles, the spin part \chi_\text{spin} can be either the antisymmetric singlet state (total spin S=0): \chi_\text{singlet} = \frac{1}{\sqrt{2}} \left( |\uparrow \downarrow \rangle - |\downarrow \uparrow \rangle \right), which pairs with a symmetric spatial wavefunction, or one of the three symmetric triplet states (total spin S=1): \chi_\text{triplet, m=1} = |\uparrow \uparrow \rangle, \quad \chi_\text{triplet, m=0} = \frac{1}{\sqrt{2}} \left( |\uparrow \downarrow \rangle + |\downarrow \uparrow \rangle \right), \quad \chi_\text{triplet, m=-1} = |\downarrow \downarrow \rangle, which pair with an antisymmetric spatial wavefunction. These spin configurations maintain the required total antisymmetry.

Exchange Energy Contribution

In the perturbative treatment of two interacting identical fermions, such as electrons, the total is given by H = H_0 + V, where H_0 = h(1) + h(2) is the sum of single-particle operators and V is the two-body interaction V(1,2) = 1/| \mathbf{r}_1 - \mathbf{r}_2 |. Assuming the electrons occupy single-particle orbitals \phi_a and \phi_b that are approximate solutions to the single-particle h \phi = \varepsilon \phi with the same eigenvalue \varepsilon for simplicity, the unperturbed energy is E_0 = 2\varepsilon. The correction to the energy is the expectation value of V computed using the properly symmetrized spatial wavefunctions required by the antisymmetry of the total fermionic wavefunction. The normalized spatial wavefunctions are the symmetric combination for the spin-singlet state (total spin S = 0) and the antisymmetric combination for the spin-triplet state (total spin S = 1): \Psi_+(1,2) = \frac{\phi_a(1)\phi_b(2) + \phi_a(2)\phi_b(1)}{\sqrt{2(1 + S^2)}} \Psi_-(1,2) = \frac{\phi_a(1)\phi_b(2) - \phi_a(2)\phi_b(1)}{\sqrt{2(1 - S^2)}} where S = \int \phi_a^*(\mathbf{r}) \phi_b(\mathbf{r}) \, d\mathbf{r} is the overlap between the orbitals. The expectation value \langle \Psi_\pm | V | \Psi_\pm \rangle evaluates to E_\pm = E_0 + \frac{J \pm K}{1 \pm S^2}, where the Coulomb (or direct) integral J and the exchange integral K are J = \iint |\phi_a(1)|^2 \, V(1,2) \, |\phi_b(2)|^2 \, dV_1 \, dV_2, K = \iint \phi_a^*(1) \phi_b(1) \, V(1,2) \, \phi_b^*(2) \phi_a(2) \, dV_1 \, dV_2. Both J and K are positive for like-charged particles and repulsive V, with J > K in typical cases. The term proportional to K constitutes the exchange energy contribution, which originates entirely from the interference between the direct and exchanged configurations in the antisymmetrized wavefunction and has no classical analog. When the orbitals are orthogonal (S = 0), the expressions simplify to E_\pm = E_0 + J \pm [K](/page/K). In this case, the exchange term -K in the antisymmetric spatial state (\Psi_-, triplet) reduces the energy relative to the direct repulsion J, acting as an effective attraction that favors parallel spins in accordance with Hund's rule. For the symmetric spatial state (\Psi_+, ), the +K term increases the repulsion. This exchange contribution is independent of and arises purely from the spatial antisymmetry imposed by the ; it explains the energy splitting between and triplet states, such as the $1s2s configuration in the , where the triplet lies $2K \approx 2.4 eV below the (experimentally \approx 0.8 eV after higher-order corrections). In the general non-orthogonal case (S \neq 0), the denominators $1 \pm S^2 further modulate the exchange effect, with overlap enhancing the magnitude of the contribution in molecular systems like H_2.

Historical Development

Heisenberg's Initial Insights

In 1926, applied the newly developed to the , addressing longstanding difficulties in treating multi-electron systems within . In his paper "Über die Spektra von Atomsystemen mit einem Kern und mehreren Elektronen," he analyzed the interaction between the two equivalent electrons, introducing the concept of quantum-mechanical resonance as a means to account for the observed and triplet spectra of para- and ortho-helium. This resonance arose from the indistinguishability of the electrons, where their wavefunctions overlapped, leading to a of states that split the energy levels without any classical counterpart, such as orbiting or direct repulsion. described this as a periodic "exchange of positions" (Platzwechsel) between the electrons, a purely quantum effect that lowered the energy of the symmetric state relative to the antisymmetric one, providing the first quantitative insight into electron correlation in atoms. Building on this, Heisenberg's subsequent work from 1927 to 1929 extended the idea to molecular systems, particularly the hydrogen molecular ion (H₂⁺), demonstrating how could act as a binding mechanism. In "Mehrkörperprobleme und Resonanz in der Quantenmechanik II," he considered the superposition of electronic states where the single could be associated with either , resulting in symmetric and antisymmetric combinations that yielded attractive and repulsive potentials, respectively. The symmetric produced a lower , enabling molecular through the overlap of wavefunctions centered on separated nuclei—a that resolved issues in for identical particles in the post-matrix mechanics era. This conceptual shift emphasized that the stemmed from the symmetry requirements of , rather than classical forces, marking the origin of as a quantum phenomenon in . Heisenberg's initial insights were groundbreaking yet limited, as his early formulations neglected electron spin, which was only recently proposed by Uhlenbeck and Goudsmit in and not fully integrated until later refinements. Without spin, his treatment relied on spatial symmetry alone to enforce the , leading to approximate calculations that captured the qualitative splitting but required subsequent adjustments for precise spectral predictions. These developments occurred amid the rapid evolution of , solving key puzzles in identical particle interactions while laying the groundwork for understanding quantum statistics in complex systems.

Extensions to Bonding and Nuclei

Following Heisenberg's initial insights into atomic resonance, the concept of exchange interactions was rapidly extended to molecular bonding and in the late and early . In 1926, incorporated exchange effects into for identical particles. This laid groundwork for understanding covalent bonds as arising from symmetric or antisymmetric combinations of atomic orbitals. The seminal application came in 1927 with Walter Heitler and Fritz London's valence bond treatment of the H₂ molecule, where the exchange integral in the singlet state (antisymmetric spin, symmetric spatial wavefunction) provides the dominant attractive contribution, explaining the stability of the covalent bond with a binding energy of approximately 4.7 eV and equilibrium distance of 0.74 Å. Fritz London further clarified in 1928 that this exchange mechanism underpins chemical binding in general, distinguishing the bonding singlet from the repulsive triplet state and emphasizing the role of resonance in valence structures. The discovery of the neutron by in 1932 enabled analogous exchange models in , shifting focus from atomic electrons to interactions. In 1932, proposed exchange forces between protons and neutrons to explain deuteron binding, treating nucleons as identical under charge exchange while incorporating spin-dependent symmetry. independently developed a symmetric in 1933, capturing proton-neutron equivalence through what later became formalism, which yields a central force saturating at short ranges and binding the deuteron with energy 2.2 MeV. By the mid-1930s, and collaborators integrated these ideas into early nuclear shell models, using exchange to enforce the Pauli principle among nucleons and explain saturation—preventing indefinite binding growth—consistent with observed like 2, 8, and 20 in stable isotopes. This evolution reframed Heisenberg's atomic "resonance" as "exchange force" in nuclear contexts, explicitly distinguishing it from later quantum field-theoretic virtual particle exchanges, as it relies on wavefunction symmetrization rather than mediation.

Applications

Atomic and Molecular Bonding

In , covalent bonds form through the overlap of atomic orbitals from adjacent atoms, leading to an exchange attraction that stabilizes the molecule when the electrons occupy a symmetric spatial wavefunction. This arises from the quantum mechanical indistinguishability of electrons, resulting in a lowering of the system's due to the symmetric combination of atomic orbitals. A classic example is the hydrogen molecule (H₂) in its , where the spin state allows the two 1s orbitals to overlap symmetrically, producing an that binds the atoms with a of approximately 104 kcal/, significantly below the of separated atoms. In contrast, the features an antisymmetric spatial wavefunction, leading to repulsion and no stable bond, as the term becomes positive and destabilizing. In polyatomic molecules, extends this concept through , where atomic orbitals mix to form hybrid orbitals optimized for maximum overlap with neighboring atoms, enhancing the exchange attraction in directional bonds, as seen in the tetrahedral sp³ of (CH₄). Resonance structures further incorporate exchange by allowing delocalization of electron pairs across multiple equivalent bonding configurations, such as in , where the cyclic overlap contributes to the molecule's . Unlike , which relies on classical electrostatic attractions between charged species, the exchange force in covalent is a purely quantum statistical effect stemming from wavefunction , without net charge transfer between atoms. The strength of this exchange-driven scales with the orbital overlap integral S, where greater overlap correlates with stronger binding, though the exact energy also depends on exchange integrals like J.

Magnetism in Solids

In solids, the governs magnetic ordering by favoring specific alignments of on sites, leading to collective phenomena such as and . This arises from the quantum mechanical exchange of between neighboring atoms, which lowers the system's energy when align in a manner dictated by the . The effective description of these interactions in systems is captured by the Heisenberg Hamiltonian, given by H = -\sum_{\langle i,j \rangle} 2J_{ij} \mathbf{S}_i \cdot \mathbf{S}_j, where the sum is over nearest-neighbor pairs \langle i,j \rangle, \mathbf{S}_i and \mathbf{S}_j are the spin operators at sites i and j, and J_{ij} is the exchange constant. When J_{ij} > 0, parallel spin alignments are energetically favored, resulting in ferromagnetism; conversely, J_{ij} < 0 promotes antiparallel alignments, yielding antiferromagnetism. Direct is the primary in many metallic and insulating , originating from the short-range spatial overlap of orbitals on neighboring atoms, which allows direct and repulsion effects. This is particularly strong in transition metals like iron, where partially filled d-orbitals enable significant overlap, stabilizing ferromagnetic order. In insulators, direct can still occur but is often weaker due to larger interatomic distances. For longer-range interactions, dominates in Mott insulators, such as transition metal oxides, where direct overlap is minimal but virtual electron hopping through intervening non-magnetic s (e.g., oxygen) mediates the . In these systems, an electron from one magnetic virtually transfers to the anion and then to a neighboring magnetic , effectively via second-order ; this typically favors antiferromagnetic alignment to minimize costs. A classic example is (MnO), where oxygen-mediated between Mn^{2+} ions (with S=5/2) establishes antiferromagnetic order below the Néel temperature of 116 K. In ferromagnetic iron, direct yields a positive J of approximately 0.1 , contributing to a high of 1043 K, above which thermal disorder disrupts the parallel spin alignment. This contrasts with MnO's antiferromagnetic (J ≈ -4 meV), which enforces alternating spins in a type-II structure. Fundamentally, these preferences stem from the : the antisymmetric wavefunction for fermions requires spin alignment (parallel for ) to allow symmetric spatial parts, thereby reducing through delocalization while avoiding double occupancy.

Nuclear Structure

In , the exchange force arises from the indistinguishability of protons and neutrons under the SU(2) symmetry, treating them as two states of a isodoublet. This symmetry implies that the total wavefunction of must be antisymmetric under particle to obey the . In the deuteron, the lightest bound nucleus consisting of one proton and one neutron, the has total T=0 (antisymmetric in isospin) and S=1 (symmetric in spin), with a dominant S-wave spatial component that is symmetric. This configuration allows an attractive , contributing to the observed of approximately 2.225 MeV. Nuclear potential models incorporate exchange effects to describe interactions, often using a Yukawa form modified by and dependencies. A common phenomenological representation is V(r) = V_{\text{central}}(r) + V_{\text{exchange}}(r) P_{\sigma} P_{\tau}, where P_{\sigma} and P_{\tau} are the -exchange and isospin-exchange operators, respectively, which capture the statistical effects of antisymmetrization. These terms arise naturally in meson-exchange theories, such as one-pion exchange, where the nature of the leads to - and isospin-dependent attractions or repulsions depending on the state symmetry. For instance, in the deuteron, the P_{\tau} operator favors the T=0 channel, enhancing binding in the -triplet state while the T=1 singlet remains unbound. The nuclear shell model relies on antisymmetrized multi-nucleon wavefunctions to enforce the Pauli principle, restricting nucleons to distinct single-particle orbitals and explaining the stability of nuclei with " of protons or neutrons (2, 8, 20, 28, 50, 82, 126). These numbers correspond to completed shells in a mean-field potential with strong spin-orbit coupling, as proposed by and J. Hans D. Jensen, where exchange interactions ensure only Pauli-allowed configurations contribute to the ground state energy. This antisymmetrization prevents overlap of identical states, leading to closed-shell nuclei with enhanced binding and low excitation energies, such as ^{4}\text{He}, ^{16}\text{O}, and ^{208}\text{Pb}. Exchange repulsion plays a crucial role in nuclear saturation, balancing the attractive central force to prevent collapse under increasing . Werner Heisenberg's early model of nuclear forces demonstrated that the antisymmetric nature of the wavefunction introduces a short-range repulsion, limiting the number of s that can bind closely and yielding a near-constant per (~8 MeV) across heavy nuclei. This mechanism ensures nuclear saturates at a density of about 0.17 s per fm³, as the term grows with the number of pairs while attraction does not scale indefinitely. A representative example is the (^{4}\text{He}), the of which features a fully symmetric spatial wavefunction in the lowest S-shell, paired with a totally antisymmetric spin-isospin component (S=0, T=0) to maintain overall antisymmetry for the four fermions. This configuration, akin to a spin-isospin singlet under SU(4) symmetry, maximizes attraction while respecting Pauli exclusion, resulting in exceptional stability with a of 28.3 MeV.

References

  1. [1]
    Exchange Forces - HyperPhysics Concepts
    All four of the fundamental forces involve the exchange of one or more particles. Even the underlying color force which is presumed to hold the quarks together.Missing: definition | Show results with:definition
  2. [2]
    [2402.00191] The development of the concept of exchange forces in ...
    Jan 31, 2024 · The onset and the development of the concept of exchange force in quantum physics are historically reconstructed, starting from Heisenberg's seminal ...
  3. [3]
    Nuclear Forces - UCLA Physics & Astronomy
    (For an ordinary force, it is essentially the same ,i.e. attractive in odd as in even states, while a space exchange force has opposite sign, i.e. repulsive, ...
  4. [4]
    [PDF] Hideki Yukawa - Nobel Lecture
    The meson theory started from the extension of the concept of the field of force so as to include the nuclear forces in addition to the gravitational and.
  5. [5]
    [PDF] Identical Particles
    Understanding the quantum mechanics of systems of identical particles requires a new postulate, the symmetrization postulate, in addition to the measurement ...
  6. [6]
    [PDF] Pauli_spin_statistics_1940.pdf - Fisica Fundamental
    In the following paper we conclude for the relativistically invariant wave equation for free particles: From postulate (I), according to which the energy must ...
  7. [7]
    [PDF] Chapter 4: Identical Particles
    Jan 4, 2024 · For instance, photons are spin-1 bosons, and electrons are spin-. 1/2 fermions. Though the spin-statistics theorem is simple to state, its proof ...
  8. [8]
    [PDF] Chapter 8 - Identical Particles - MIT OpenCourseWare
    May 8, 2022 · Spin-statistics theorem from Quantum Field Theory shows that bosons are particles of integer spins (0, 1, 2, . . . ) while fermions are ...
  9. [9]
    [PDF] Identical Particles
    The sign of σ is of crucial importance for the properties of the particles. The properties of bosons and fermions are generally speaking radically different.
  10. [10]
    [1902.00499] Pauli Exclusion Principle and its theoretical foundation
    Jan 29, 2019 · It asserts that particles with half-integer spin (fermions) are described by antisymmetric wave functions, and particles with integer spin (bosons) are ...
  11. [11]
    January 1925: Wolfgang Pauli announces the exclusion principle
    Pauli originally applied the exclusion principle to explain electrons in atoms, but later it was extended to any system of fermions, which have half integer ...
  12. [12]
    Spin effects and the Pauli principle in semiclassical electron dynamics
    Mar 3, 2014 · This demonstrates the Pauli principle: Two electrons in a triplet cannot come as close to each other as two electrons in a singlet.<|separator|>
  13. [13]
    Spin-Resolved Mapping of Spin Contribution to Exchange ...
    Feb 26, 2010 · The Pauli principle demands that two electrons with parallel spins can not be at the same location, while the Coulomb correlation will force ...Missing: exclusion | Show results with:exclusion
  14. [14]
    [PDF] Wolfgang Pauli - Nobel Lecture
    Already in my original paper I stressed the circumstance that I was unable to give a logical reason for the exclusion principle or to deduce it from more. Page ...
  15. [15]
    [PDF] Quantum Mechanics, The Stability of Matter and Quantum ... - arXiv
    Jan 4, 2004 · If Coulomb forces are involved, then the Pauli exclusion principle is essential. (This means that the L2 functions of N vari- ables, Ψ(x1, x2,..
  16. [16]
    [PDF] Chapter 4: Identical Particles - MIT OpenCourseWare
    As an application, a pair of electrons must have either a symmetric spatial wavefunction and an antisymmetric spin wavefunction (i.e. singlet), or vice versa, ...
  17. [17]
    [PDF] Identical particles - Advanced quantum mechanics (20104301)
    If ψ1 and ψ2 are the respective normalized wave functions, the approximately normalized antisymmetric wave function is given by ψ(ξ1, ξ2) = 1. √. 2. (ψ1(ξ1)ψ2 ...
  18. [18]
    [PDF] Identical Particles - TCM
    It is evident that the spin singlet wavefunction is antisymmetric under the exchange of two particles, while the spin triplet wavefunction is symmetric. For a ...
  19. [19]
    [PDF] TWO ELECTRONS: EXCITED STATES
    In the last lecture, we learned that the independent particle model gives a reasonable description of the ground state energy of the Helium atom.
  20. [20]
    [PDF] Many-Body Problems III (Molecular Physics) - 221B Lecture Notes
    The potential energy is actually lower for the anti-symmetric one because the anti-symmetry min- imizes the electron-electron repulsion. But the difference in ...
  21. [21]
  22. [22]
  23. [23]
    On the theory of quantum mechanics - Journals
    The new mechanics of the atom introduced by Heisenberg may be based on the assumption that the variables that describe a dynamical system do not obey the ...
  24. [24]
    Valence Bond Theory—Its Birth, Struggles with Molecular Orbital ...
    Mar 15, 2021 · ... exchange energy. In modern terms and pictorially, the bonding in H2 can be accounted for by the wave function drawn in Scheme 1. This wave ...Missing: H2 | Show results with:H2
  25. [25]
    [PDF] THE NATURE OF THE CHEMICAL BOND MMI ij m
    Mar 1, 2025 · A detailed discussion of the valence-bond theory of the electronic structure of metals and intermetallic compounds is also presented.
  26. [26]
    Quantum electrodynamics and the origins of the exchange, dipole ...
    Understanding the origin of ferromagnetism was one of the success stories of applying quantum mechanics to solid-state systems. Classically, magnetic moments ...
  27. [27]
    Theory of Direct Exchange in Ferromagnetism | Phys. Rev.
    An extensive investigation is presented on the role of direct exchange as the mechanism responsible for ferromagnetism.<|control11|><|separator|>
  28. [28]
    Antiferromagnetism. Theory of Superexchange Interaction | Phys. Rev.
    In this paper the general formalism of Kramers indicating the existence of superexchange interaction has been reduced, under simplifying assumptions.Missing: MnO | Show results with:MnO
  29. [29]
    [0904.1297] Spin--orbital physics in transition metal oxides - arXiv
    Apr 8, 2009 · We present the main features of the spin-orbital superexchange which describes the magnetic and optical properties of Mott insulators with ...
  30. [30]
    Verification of Anderson Superexchange in MnO via Magnetic Pair ...
    This magnetic structure consists of ferromagnetically aligned spins within common (111) sheets and antiferromagnetic coupling between adjacent sheets. All fits ...
  31. [31]
    Flat-Band Ferromagnetism as a Pauli-Correlated Percolation Problem
    The interplay of the Coulomb interaction with the Pauli principle was already recognized by Heisenberg [1] to give rise to a ferromagnetic exchange ...
  32. [32]
    [PDF] LECTURE 3 – SU(2) - Particle Physics Department
    As far as strong interactions are concerned, the proton and neutron are indistinguishable and hence in. QM, they are interchangeable. ... Consequently there is an ...
  33. [33]
    [PDF] Nucleon-Nucleon Interaction, Deuteron - UMD Physics
    6.4 Deuteron Structure. The deuteron is the only bound state of 2 nucleons, with isospin T = 0, spin-parity Jπ = 1+, and binding energy EB=2.225 MeV. For two ...
  34. [34]
    One Pion Exchange Potential.
    Let us construct the potential arising from one-pion exchange (OPE), it arises from the tree-level graphs with four external nucleons and an off-shell pion.
  35. [35]
    [PDF] Effective Field Theories • Yukawa potential • Nucleon ... - UBC Physics
    Yukawa model of nuclear forces 1934 (Frauenfelder and Henley). • Scalar ... • A detailed model of pion exchange: 'One Pion Exchange Potential' (OPEP):. V ...
  36. [36]
    [PDF] Mayer–Jensen Shell Model and Magic Numbers
    Nuclear shell model, magic num- bers, spin-orbit coupling. M any nuclearm odelshave been putforw ard since. 1932. A m ong them the collective m odel pro ...
  37. [37]
    [PDF] The Nuclear Shell Model - High Energy Physics |
    Magic. Numbers. Page 12. 326. Predictions of the Nuclear Shell Model. (1) MAGIC NUMBERS : The Shell Model successfully predicts the origin of the magic numbers.
  38. [38]
    [PDF] PoS(EMC2006)002 - SISSA
    Oct 6, 2006 · Majorana and Heisenberg must have discussed the exchange theory of nuclear forces during those first days. Heisenberg's third paper [4] was not.
  39. [39]
    [PDF] Lecture 7 1.1. If we add two vectors of lengths r and r the sum can ...
    The α particle is a spin zero and isospin zero state; it can be thought of as a bound state of four nucleons. The αparticle is the nu- cleus of the abundant ...