Fact-checked by Grok 2 weeks ago

Many-body problem

The many-body problem in physics refers to the challenge of predicting the collective behavior and properties of systems composed of a large number of interacting particles, where exact analytical solutions are generally impossible beyond a few bodies due to the exponential increase in complexity with the number of particles. This problem spans classical and quantum regimes, with the classical formulation known as the , which describes the motion of n point masses under mutual gravitational attraction according to and universal gravitation. In the quantum case, it involves solving the many-particle for interacting particles, such as electrons in solids or nucleons in atomic nuclei, leading to emergent phenomena like and . Historically, the classical many-body problem originated in during the , with formulating the foundational equations in his (1687), building on Johannes Kepler's empirical laws of planetary motion. Henri Poincaré's work in the late demonstrated the non-integrability and chaotic nature of the problem for n \geq 3, influencing the development of and . The quantum many-body problem gained prominence in the mid-20th century, driven by discoveries of macroscopic quantum effects like in (1937) and in metals (1911), which required new theoretical frameworks to explain collective excitations in interacting particle systems. Key challenges in the many-body problem include the "curse of dimensionality," where the configuration space grows factorially with particle number, rendering exact solutions infeasible and necessitating approximations such as mean-field theories, perturbation methods, or numerical simulations like Monte Carlo or density functional theory. In classical systems, conserved quantities like energy, linear momentum, angular momentum, and center-of-mass motion provide 10 integrals of motion, but these are insufficient to fully specify trajectories for n > 2, leading to reliance on statistical mechanics for large N. Quantum treatments often employ second quantization and Green's functions to handle indistinguishable particles and correlations, with tools like Feynman diagrams facilitating calculations of interaction effects. The many-body problem is central to numerous fields: in , it underpins the study of electron gases, phase transitions, and ; in , it models the structure and reactions of atomic nuclei; and in , it simulates gravitational dynamics in star clusters or galaxies via N-body codes. Recent advances, such as Bose-Einstein condensation in ultracold atomic gases (achieved in 1995), have enabled experimental probes of many-body effects, while computational methods like continue to push boundaries in simulating strongly correlated systems. Overall, solving aspects of the many-body problem has revolutionized our understanding of complex systems, from everyday materials to the universe's largest structures.

Overview

Definition and Scope

The many-body problem refers to the challenge of predicting the collective behavior and properties of a composed of more than two interacting particles, where exact analytical solutions become impossible for N \geq 3 due to the exponential increase in with the number of particles. This problem originated in efforts to understand interactions in but extends broadly across physics to describe complex systems in both microscopic and macroscopic scales. Its scope spans classical and quantum regimes: in the classical domain, it focuses on deterministic trajectories governed by Newtonian laws, whereas in , it deals with probabilistic descriptions via wavefunctions that capture entangled states and correlations among particles. A key distinction lies in the transition from few-body to many-body systems. The is analytically solvable by reducing it to an equivalent one-body problem through center-of-mass separation, allowing exact orbits or states to be determined. In contrast, for N > 2, the coupled interactions lead to non-integrable dynamics in the classical case and an exponentially large in the quantum case, necessitating approximate methods to extract meaningful predictions about system properties such as energy levels, phase transitions, or transport behaviors. Mathematically, the classical many-body problem is formulated using the H = \sum_{i=1}^N \frac{\mathbf{p}_i^2}{2m} + \sum_{1 \leq i < j \leq N} V(\mathbf{r}_i - \mathbf{r}_j), where \mathbf{p}_i and \mathbf{r}_i are the momentum and position of the i-th particle, m is the mass, and V is the pairwise interaction potential; Hamilton's equations then yield the equations of motion. In the quantum regime, the system's evolution is governed by the time-dependent Schrödinger equation i \hbar \frac{\partial \psi}{\partial t} = \hat{H} \psi, with the many-body \hat{H} incorporating kinetic energy operators -\frac{\hbar^2}{2m} \nabla_i^2 for each particle and the same interaction potential V(\mathbf{r}_i - \mathbf{r}_j), resulting in a high-dimensional wavefunction \psi(\mathbf{r}_1, \dots, \mathbf{r}_N, t).

Historical Context

While the many-body problem in classical mechanics was first formulated in the 17th century, fundamental limitations in solving systems of interacting particles were revealed in the late 19th century. In the 1890s, Henri Poincaré analyzed the three-body problem within Newtonian gravity, demonstrating through qualitative methods that it lacks a general analytic solution and exhibits chaotic behavior, thereby establishing non-integrability for broader N-body systems. This work, detailed in his treatise Les Méthodes Nouvelles de la Mécanique Céleste, shifted focus from exact integrability to qualitative dynamics, laying groundwork for understanding complex interactions beyond two bodies. A pivotal milestone came with the introduction of statistical mechanics, which addressed the intractability of exact solutions by averaging over ensembles of many-body configurations. In 1902, Josiah Willard Gibbs formalized this approach in Elementary Principles in Statistical Mechanics, providing a probabilistic framework to describe thermodynamic properties of large particle assemblies without solving individual trajectories. This method proved essential for treating classical many-body systems in gases and liquids, bridging microscopic interactions to macroscopic observables. The advent of quantum mechanics in the 1920s and 1930s transformed the many-body problem, extending classical challenges to wave-like particle behaviors and indistinguishability. Werner Heisenberg's matrix mechanics, developed in 1925, enabled initial applications to atomic systems, while his subsequent work in the mid-1920s explored multi-electron configurations in atoms like helium, highlighting correlation effects. Paul Dirac advanced this further in 1929 with his antisymmetrized wavefunction for many-electron systems, incorporating exchange interactions to align theory with atomic spectra. In the 1930s, Hans Bethe applied these quantum ideas to nuclear physics, developing models for nucleon interactions and shell structures that captured many-body effects in atomic nuclei. Post-World War II research intensified focus on condensed matter, where many-body quantum effects dominate material properties. Felix Bloch and contemporaries like John Bardeen shifted emphasis to solid-state systems, building on Bloch's earlier wave theorems to investigate electron correlations in crystals and metals during the late 1940s and 1950s. Concurrently, Richard Feynman's path integral formulation, introduced in 1948, offered a new quantum tool for summing over all possible particle histories, facilitating treatments of interacting many-body states in both non-relativistic and field-theoretic contexts. These developments marked a transition from isolated atomic studies to unified frameworks for diverse physical regimes.

Classical Many-Body Problem

Newtonian Formulation

In the Newtonian formulation of the classical many-body problem, a system consists of N particles, each with position vector \mathbf{r}_i(t) and mass m_i, interacting through pairwise forces derived from a potential energy function V(\{\mathbf{r}_j\}). The motion is governed by Newton's second law applied to each particle: m_i \frac{d^2 \mathbf{r}_i}{dt^2} = -\nabla_i V(\{\mathbf{r}_j\}), where \nabla_i denotes the gradient with respect to \mathbf{r}_i, and the total force on particle i arises from the negative gradient of the potential, assuming conservative interactions. This set of $3N second-order differential equations describes the deterministic evolution of the system in three-dimensional Euclidean space, with the potential V typically depending on the relative distances between particles, such as in gravitational or electrostatic interactions. The Lagrangian formulation provides an alternative, coordinate-independent approach to derive these equations. The Lagrangian L is defined as the difference between the total kinetic energy T = \sum_{i=1}^N \frac{1}{2} m_i |\dot{\mathbf{r}}_i|^2 and the potential energy V: L = T - V. Using generalized coordinates q_j (which may include the Cartesian components of \mathbf{r}_i), the Euler-Lagrange equations yield the dynamics: \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_j} \right) - \frac{\partial L}{\partial q_j} = 0 for each j = 1, \dots, 3N. The Hamiltonian formulation, obtained via a Legendre transform, introduces conjugate momenta \mathbf{p}_i = m_i \dot{\mathbf{r}}_i, resulting in the total energy as the Hamiltonian: H = \sum_{i=1}^N \frac{|\mathbf{p}_i|^2}{2 m_i} + V(\{\mathbf{r}_j\}). Hamilton's equations then govern the evolution: \dot{\mathbf{r}}_i = \frac{\partial H}{\partial \mathbf{p}_i} and \dot{\mathbf{p}}_i = -\frac{\partial H}{\partial \mathbf{r}_i}, preserving the symplectic structure of phase space. For few-particle systems, the many-body problem simplifies significantly. In the two-body case, the equations decouple into center-of-mass motion \mathbf{R} = \sum m_i \mathbf{r}_i / M (with total mass M = \sum m_i) and relative motion \mathbf{r} = \mathbf{r}_2 - \mathbf{r}_1, reducible to an effective one-body problem with reduced mass \mu = m_1 m_2 / (m_1 + m_2) and potential V(|\mathbf{r}|). This yields closed-form solutions for central potentials, such as elliptical orbits in gravity. For three or more bodies, no such general reduction exists, and the full $3N-dimensional system must be solved, highlighting the onset of complexity even at small N. In isolated systems with no external forces or torques and time-independent potentials, several quantities are conserved. Linear momentum \mathbf{P} = \sum m_i \dot{\mathbf{r}}_i is preserved due to translational invariance, angular momentum \mathbf{L} = \sum \mathbf{r}_i \times \mathbf{p}_i due to rotational invariance, and total energy E = T + V (or H) due to time independence. These ten integrals (six for center-of-mass motion, three for angular momentum, one for energy) constrain the dynamics but are insufficient for N > 2 to yield complete integrability.

Integrability and Non-Integrability

In classical , a system with N , possessing a $2N-dimensional [phase space](/page/Phase_space), is deemed completely integrable if it admits N$ independent integrals of motion that are in involution with respect to the , allowing the motion to be solved via and quadratures. This criterion, formalized by the Liouville-Arnold theorem, ensures that the foliates into invariant tori on which trajectories are quasi-periodic, providing a complete qualitative and quantitative description of the dynamics. The two-body problem under inverse-square gravitation, equivalent to the Kepler problem, exemplifies integrability, as it possesses the required independent conserved quantities: total energy, angular momentum vector, and the Laplace-Runge-Lenz vector, enabling exact elliptic solutions. In contrast, the general N-body problem for N \geq 3 is non-integrable, as established by the Bruns-Poincaré theorem, which proves that no additional independent algebraic integrals exist beyond the ten classical ones (energy, total momentum, center-of-mass position, and angular momentum). Henri Poincaré's investigations into the three-body problem further illuminated this non-integrability by revealing the absence of uniform first integrals beyond the known ones, laying groundwork for qualitative analysis of irregular orbits. For systems near integrability, such as the N-body problem with small perturbations from exactly solvable cases, the Kolmogorov-Arnold-Moser (KAM) theorem demonstrates that most invariant tori persist, resulting in quasi-integrable behavior where a positive measure of retains quasi-periodic motion, though interspersed with chaotic regions. Chaos emerges prominently in non-integrable many-body dynamics, quantified by , which measure the exponential divergence of nearby trajectories; the largest positive Lyapunov exponent \lambda_1 > 0 indicates local instability, with the spectrum scaling in a manner that reflects the system's approach to thermalization in the classical limit. These properties imply that exact analytical solutions are unattainable for the classical N-body problem when N > 2, necessitating numerical integration of the equations of motion to study individual trajectories, while for large N, the prevalence of ergodic behavior on energy surfaces underpins the development of statistical mechanics to describe ensemble averages rather than precise orbits.

Quantum Many-Body Problem

Wavefunction Description

In quantum mechanics, the many-body wavefunction \psi(\mathbf{r}_1, \dots, \mathbf{r}_N, t) provides a complete description of the state of a system consisting of N identical particles, where \mathbf{r}_i denotes the position vector of the i-th particle and t is time. This function evolves according to the time-dependent Schrödinger equation and encodes all observable properties of the system through its probability density |\psi|^2, which gives the likelihood of finding the particles at specific positions. For identical particles, the principle of indistinguishability requires the wavefunction to transform in a specific way under particle exchange: it must be totally symmetric for bosons (integer-spin particles like photons) and totally antisymmetric for fermions (half-integer-spin particles like electrons). The antisymmetric form for fermions is commonly expressed as a Slater determinant, a mathematical construct that automatically enforces the required exchange symmetry while incorporating single-particle orbitals. The domain of the many-body wavefunction is the $3N-dimensional configuration space, formed by the Cartesian coordinates of all N particles, which introduces an exponential scaling in the as N increases. This high dimensionality underlies the computational intractability of exact solutions for large systems, as the wavefunction must be specified over a vast whose volume grows exponentially with N. In contrast to the classical many-body problem, where trajectories are determined in 3D , the quantum configuration space captures superposition and delocalization effects inherent to wave-like behavior. For stationary states, the time-independent governs the energy eigenfunctions: \hat{H} \psi = E \psi, where \hat{H} is the many-body . Explicitly, for non-relativistic particles of equal m, \hat{H} = -\frac{\hbar^2}{2m} \sum_{i=1}^N \nabla_i^2 + \sum_{i < j}^N V(|\mathbf{r}_i - \mathbf{r}_j|) + \sum_{i=1}^N V_{\text{ext}}(\mathbf{r}_i), with the first term representing the kinetic energy (via the Laplacian \nabla_i^2 acting on the i-th particle's coordinates), the second term the two-body interaction potential V, and the third any external potentials V_{\text{ext}}. This formulation generalizes the single-particle to account for interactions, highlighting the challenge of solving for \psi in high dimensions. A key feature of many-body wavefunctions is quantum entanglement, where the state of the entire system cannot be separated into independent states for subsystems, leading to non-local correlations. For few-particle systems, such as two qubits, entanglement is exemplified by Bell states like \frac{1}{\sqrt{2}} (|01\rangle - |10\rangle), which violate classical intuitions about separability. In many-body contexts, entanglement is quantified using the von Neumann entropy S = -\operatorname{Tr}(\rho \log_2 \rho) of the reduced density matrix \rho for a subsystem, providing a measure of the entanglement across a bipartition; for ground states of local Hamiltonians, this entropy often scales with the boundary area rather than the volume of the subsystem, known as an area law.

Second Quantization Framework

Second quantization provides a powerful formalism for describing quantum many-body systems, particularly those involving identical particles, by reformulating the theory in terms of creation and annihilation operators acting on a Fock space. This approach, originally developed in the late 1920s and early 1930s, shifts the focus from explicit wavefunctions of fixed particle number to an operator algebra that naturally accommodates variable particle numbers and particle statistics. The Fock space serves as the Hilbert space for second quantization, constructed as a direct sum of N-particle Hilbert spaces for N = 0, 1, 2, \dots, with the vacuum state |0\rangle as the N=0 sector. States in Fock space are built from single-particle basis states labeled by a mode index k, using occupation number representations where the basis vectors are specified by the number of particles n_k in each mode, ensuring symmetrization or antisymmetrization according to particle type. This structure allows for a unified treatment of systems with arbitrary particle numbers, avoiding the need to specify N explicitly in advance. Central to the formalism are the creation operator \hat{a}^\dagger_k and annihilation operator \hat{a}_k for mode k. For bosons, these satisfy the commutation relations [\hat{a}_k, \hat{a}^\dagger_l] = \delta_{kl} and [\hat{a}_k, \hat{a}_l] = [\hat{a}^\dagger_k, \hat{a}^\dagger_l] = 0, generating symmetric states from the vacuum via |n_k\rangle = \frac{(\hat{a}^\dagger_k)^{n_k}}{\sqrt{n_k!}} |0\rangle. For fermions, anticommutation relations apply: \{\hat{a}_k, \hat{a}^\dagger_l\} = \delta_{kl} and \{\hat{a}_k, \hat{a}_l\} = \{\hat{a}^\dagger_k, \hat{a}^\dagger_l\} = 0, yielding antisymmetric states with occupation numbers restricted to n_k = 0 or $1, as \hat{a}_k |1_k\rangle = |0\rangleand(\hat{a}^\dagger_k)^2 |0\rangle = 0$. These relations enforce Bose-Einstein or Fermi-Dirac statistics inherently, without imposing symmetry constraints manually on wavefunctions. In second quantization, the many-body Hamiltonian is expressed using these operators. The non-interacting part is \hat{H}_0 = \sum_k \varepsilon_k \hat{a}^\dagger_k \hat{a}_k, where \varepsilon_k is the single-particle energy for mode k, and \hat{n}_k = \hat{a}^\dagger_k \hat{a}_k counts particles in that mode. Interactions are captured by the two-body term \hat{H}_\text{int} = \frac{1}{2} \sum_{k,l,m,n} V_{klmn} \hat{a}^\dagger_k \hat{a}^\dagger_l \hat{a}_m \hat{a}_n, with V_{klmn} the matrix elements of the interaction potential, allowing for general scattering processes while preserving particle statistics through operator ordering. This operator form facilitates the treatment of both finite and infinite systems, such as in condensed matter or quantum field theory. The advantages of second quantization include its ability to handle systems with fluctuating particle numbers, as in grand canonical ensembles, and to simplify calculations involving identical particles by automatically incorporating exchange effects via the commutation relations. Unlike first quantization, where wavefunction antisymmetry must be enforced explicitly for fermions, this framework embeds such requirements in the algebra, streamlining derivations for correlation functions and response properties.

Core Challenges

Computational Intractability

The quantum many-body problem exhibits profound computational intractability due to the exponential growth of the Hilbert space dimension with the number of particles. For a system of N indistinguishable particles, each occupying one of d single-particle states, the dimension of the full Hilbert space scales as \binom{d}{N} \approx d^N / N! in the fermionic case, leading to storage requirements and computational time that grow exponentially as \exp(N). This exponential scaling arises because the many-body wavefunction must describe all possible configurations of the particles, entangling them across the vast state space. This phenomenon, known as the curse of dimensionality, renders exact solutions via direct diagonalization of the many-body infeasible for all but the smallest systems. For electronic systems, full configuration interaction methods—equivalent to exact diagonalization in a finite basis—are practically limited to molecules with fewer than 20-22 electrons in comparable orbitals, beyond which the matrix sizes exceed current computational capabilities. Quantum Monte Carlo methods offer partial mitigation by stochastically sampling the , enabling studies of larger systems in certain cases, but they are hampered by the fermion sign problem, where oscillating wavefunction phases lead to exponential variance in estimators and severely restrict applicability to fermionic systems at low temperatures or away from half-filling. In classical many-body systems, a analogous challenge emerges from the growth of phase space volume, which, informed by Heisenberg's uncertainty principle, is discretized into cells of minimum volume h^{3N} for N particles, where h is Planck's constant. However, the quantum case is more severe due to the non-separable nature of entangled states, which cannot be decomposed into independent single-particle trajectories, amplifying the resource demands beyond classical statistical sampling. From a complexity theory perspective, determining the ground state energy of a local many-body Hamiltonian to within additive error is QMA-complete, as established in the early 2000s, implying that no efficient classical algorithm exists unless quantum Merlin-Arthur equals polynomial time (QMA = P), a conjecture considered unlikely. This hardness holds even for Hamiltonians with 5-local terms, underscoring the intrinsic difficulty of the problem for realistic, sparse interactions in quantum systems.

Correlation Effects

In quantum many-body systems, particularly those involving electrons, correlation effects manifest as deviations from independent-particle behavior due to interparticle interactions, primarily and . These effects prevent the accurate description of the system's ground state and excitations using simple single-determinant wavefunctions, as the positions and motions of particles are interdependent, leading to a more complex . Correlation effects are categorized into static and dynamic types based on their spatial and temporal scales. Static correlations are long-range and arise from structural near-degeneracies in the electronic configuration space, such as those captured by , which dictate the preference for high-spin, maximally symmetric states in degenerate orbitals to reduce on-site repulsion in multi-orbital systems. In contrast, dynamic correlations involve short-range, time-dependent fluctuations where electrons adjust their positions instantaneously to minimize repulsion; a key example is the Coulomb hole in the uniform electron gas, a depletion region around each electron that reflects the average avoidance due to pairwise Coulomb interactions. These correlations contribute to the exchange-correlation energy, which accounts for quantum exchange and the remaining Coulomb effects beyond classical electrostatics. In the Hartree-Fock approximation, exchange is treated exactly within the single-determinant framework, but the full exchange-correlation energy, including correlation, is approximated, with its exact form unknown and often expressed in local density approximations as E_{xc} = \int \rho(\mathbf{r}) v_{xc}(\mathbf{r}) \, d\mathbf{r}. The impact of correlation effects is profound, enabling phenomena such as , where strong on-site repulsions localize electrons and suppress conductivity, and , where attractive correlations facilitate electron pairing. These interactions are characterized by the pair correlation function g(\mathbf{r}) = \frac{\langle \rho(\mathbf{r}) \rho(0) \rangle}{\rho^2}, which quantifies the enhanced or reduced probability of finding two electrons at separation \mathbf{r} relative to a non-interacting uniform distribution, with g(\mathbf{r}) < 1 at short distances indicating repulsion-dominated behavior. For systems dominated by strong correlations, where mean-field treatments fail to capture multi-reference character, configuration interaction methods are required, expanding the wavefunction as a linear combination of multiple Slater determinants to recover the correlation energy accurately through inclusion of higher-order excitations.

Approximation Methods

Mean-Field Theories

Mean-field theories approximate the complex interactions in many-body systems by replacing them with an effective average field experienced by each particle, thereby reducing the problem to a set of independent single-particle equations that can be solved self-consistently. This approach, foundational in both classical and quantum many-body physics, simplifies computational demands while capturing essential qualitative features, such as orbital structures in atoms and molecules. Originating in the early days of quantum mechanics, these methods assume that each particle moves in a potential generated by the average distribution of all others, neglecting instantaneous correlations. The Hartree method, introduced by Douglas Hartree in 1928, represents one of the earliest mean-field approximations for multi-electron atoms. In this framework, the many-electron wavefunction is approximated as a product of single-particle orbitals \psi_j(\mathbf{r}), and each electron evolves in a self-consistent potential that includes the nuclear attraction and the averaged Coulomb repulsion from all other electrons. The Hartree potential V_{\text{H}}(\mathbf{r}) is given by V_{\text{H}}(\mathbf{r}) = \sum_{j=1}^{N-1} \int \frac{|\psi_j(\mathbf{r}')|^2}{|\mathbf{r} - \mathbf{r}'|} \, d\mathbf{r}', where the sum runs over the occupied orbitals excluding the one under consideration, and the integral computes the classical electrostatic potential due to the electron density. The orbitals are then determined by solving the resulting single-particle Schrödinger-like equations iteratively until convergence, providing a tractable way to estimate atomic energies and densities for systems like sodium. A significant advancement came with the Hartree-Fock (HF) method, independently developed by Vladimir Fock in 1930, which incorporates the quantum mechanical exchange effects arising from the indistinguishability of electrons. Unlike the Hartree approach, HF employs an antisymmetrized wavefunction in the form of a to ensure Pauli exclusion, minimizing the expectation value of the Hamiltonian variationally. This leads to effective orbitals \phi_i(\mathbf{r}) that satisfy the : \left( -\frac{\nabla^2}{2} + V_{\text{eff}}(\mathbf{r}) \right) \phi_i(\mathbf{r}) = \varepsilon_i \phi_i(\mathbf{r}), where V_{\text{eff}}(\mathbf{r}) combines the Hartree potential with a nonlocal exchange term that accounts for the antisymmetry. The resulting eigenvalues \varepsilon_i approximate single-particle energies, enabling predictions of molecular geometries and spectroscopic properties with reasonable accuracy for weakly correlated systems. Despite their utility, mean-field theories like Hartree-Fock have notable limitations, primarily the neglect of electron correlation effects beyond the mean field, which leads to systematic errors in energy calculations. For instance, HF overestimates band gaps in solids and molecules by 50-100% compared to experimental values, as the absence of dynamic correlations fails to properly screen interactions and delocalize electrons. These shortcomings are particularly evident in strongly correlated systems, where higher-order effects dominate. A simpler variant, the Thomas-Fermi model, treats the electron gas as a continuum and approximates the kinetic energy locally from the density, serving as an early mean-field tool for heavy atoms. Developed independently by Llewellyn Thomas and in 1927, it replaces orbital descriptions with a statistical density functional, yielding qualitative insights into atomic sizes and binding energies but lacking shell-structure details due to its semiclassical nature. This approach laid groundwork for later density-based methods while highlighting the trade-offs in mean-field simplifications.

Perturbation Techniques

Perturbation techniques address the many-body problem by expanding the Hamiltonian as \hat{H} = \hat{H}_0 + \lambda \hat{V}, where \hat{H}_0 is a solvable unperturbed part (often the mean-field Hamiltonian), \hat{V} is the interaction, and \lambda is a formal coupling parameter set to 1 after expansion. This approach, rooted in (RSPT), yields corrections to energies and wavefunctions as power series in \lambda. For the nth eigenstate, the energy is given by E_n = E_n^{(0)} + \lambda \langle \psi_n^{(0)} | \hat{V} | \psi_n^{(0)} \rangle + \lambda^2 \sum_{m \neq n} \frac{|\langle \psi_m^{(0)} | \hat{V} | \psi_n^{(0)} \rangle|^2}{E_n^{(0)} - E_m^{(0)}} + \mathcal{O}(\lambda^3), where E_n^{(0)} and \psi_n^{(0)} are the unperturbed energy and wavefunction. Higher-order terms involve increasingly complex contributions from virtual excitations, providing systematic improvements for weak interactions. In the many-body context, RSPT is formulated using second quantization to handle indistinguishable particles efficiently. Many-body perturbation theory (MBPT) employs Goldstone diagrams—graphical representations of Wick-contracted operator products—to enumerate and compute these terms, ensuring only connected (linked) diagrams contribute via the linked-cluster theorem. This theorem, established for fermionic systems, guarantees that energy corrections are extensive (additive for non-interacting subsystems) and avoids overcounting in the cluster expansion. Goldstone diagrams depict particle (upward) and hole (downward) lines, with interaction vertices, facilitating automated evaluation up to high orders for practical computations. A prominent application of MBPT is the GW approximation, which computes quasiparticle energies by approximating the self-energy \Sigma as the product of the one-particle Green's function G and the screened Coulomb interaction W. Starting from , the quasiparticle energy \varepsilon satisfies the Dyson equation \varepsilon = \varepsilon_{\mathrm{HF}} + \langle \Sigma(\varepsilon) - V_{\mathrm{xc}} \rangle, where V_{\mathrm{xc}} is the exchange-correlation potential; this yields accurate band structures in semiconductors and metals by accounting for screening beyond mean-field. Despite its successes, MBPT convergence is limited in strongly correlated regimes, where higher-order terms grow rapidly or diverge. In the , which captures electron correlations via on-site repulsion U, perturbation expansions around the non-interacting limit fail for strong coupling (U/t \gtrsim 3, with t the hopping), overestimating energy gaps and predicting spurious phase transitions absent in exact solutions. For instance, applied to finite induces artificial antiferromagnetic ordering at U \approx 2.4t for even site numbers, diverging from the exact paramagnetic ground state.

Numerical Approaches

Monte Carlo Methods

Monte Carlo methods provide a class of stochastic numerical techniques that leverage random sampling to evaluate high-dimensional integrals inherent in many-body quantum systems, enabling the computation of ground-state energies, correlation functions, and thermodynamic properties. These approaches are particularly suited to tackling the computational intractability of the many-body problem by approximating expectation values over vast configuration spaces without relying on deterministic expansions. By generating representative samples from probability distributions derived from trial wave functions, Monte Carlo simulations achieve controlled statistical errors, making them indispensable for studying strongly correlated systems where traditional methods falter. Variational Monte Carlo (VMC) forms the foundation of these techniques, applying the variational principle to minimize the energy expectation value for a parameterized trial wave function \psi_T(\mathbf{R}), where \mathbf{R} denotes the particle coordinates. The ground-state energy is estimated as E = \frac{\langle \psi_T | \hat{H} | \psi_T \rangle}{\langle \psi_T | \psi_T \rangle} = \int d\mathbf{R} \, |\psi_T(\mathbf{R})|^2 \frac{\hat{H} \psi_T(\mathbf{R})}{\psi_T(\mathbf{R})}, with the integral evaluated via Monte Carlo sampling from the distribution |\psi_T(\mathbf{R})|^2 using algorithms such as the . Introduced by McMillan in 1965 for the ground state of liquid helium-4, VMC efficiently handles bosonic systems and provides upper bounds to the true energy, with accuracy depending on the quality of the trial function, often incorporating to capture correlations. This method scales favorably for systems up to dozens of particles, offering insights into pair correlations and structural properties. Building on VMC, diffusion Monte Carlo (DMC) projects an initial trial wave function toward the exact ground state through imaginary-time evolution under the operator e^{-\tau (\hat{H} - E_T)}, where \tau is the projection time and E_T a trial energy shift. This evolution is simulated as a branching random walk in configuration space, where walkers diffuse according to the Green's function of the kinetic energy operator and branch based on local potential energy to sample the ground-state distribution. Originating from Green function Monte Carlo developments by Kalos in 1970 for bosonic fluids, DMC was extended to fermions by Ceperley, Chester, and Kalos in 1977 using a short-time approximation, and notably applied to the uniform electron gas by Ceperley and Alder in 1980, yielding energies accurate to within 1% of exact values for small densities. DMC surpasses VMC by relaxing trial function dependencies, providing mixed-estimator energies close to the variational upper bound. A primary limitation in fermionic systems stems from the fermion sign problem, where the antisymmetric wave function induces oscillatory signs in the integrands, leading to exponentially vanishing average signals amid growing noise as system size or inverse temperature increases. This issue, first prominently encountered in determinant-based formulations for lattice models by Hirsch in 1985, restricts unbiased DMC and VMC applications to bosonic or spin-polarized systems, or requires approximations like the fixed-node method, which constrains walker diffusion to nodal pockets of a trial to enforce antisymmetry. Without such fixes, simulations become infeasible for generic fermionic Hamiltonians due to the sign oscillations. In terms of accuracy, Monte Carlo methods yield statistically exact results for tractable systems, with the primary error arising from finite sampling and scaling as $1/\sqrt{M}, where M is the number of independent configurations; for instance, DMC achieves chemical accuracy (1 kcal/mol) for atomic and molecular systems with modest M \sim 10^4. This statistical convergence, combined with systematic biases from approximations like fixed-node (typically 1-5% for energies in solids), positions these techniques as benchmarks for small-to-medium systems, exact in the infinite-M limit for bosons. Recent advances as of 2025 include hybrid quantum-classical algorithms integrating quantum computing with to mitigate the sign problem and extend simulations to larger systems, such as full configuration interaction QMC on quantum hardware, demonstrating improved efficiency for strongly correlated materials.

Density Functional Theory

reformulates the many-body problem for interacting electrons in an external potential by focusing on the ground-state electron density \rho(\mathbf{r}) as the central variable. The foundational , proved in 1964, assert that the ground-state density uniquely determines the external potential and thus all ground-state properties of the system, a result known as the invertibility of the density-potential mapping. The theorems also establish that the total energy is a universal functional of the density, E[\rho], minimized by the true ground-state density among all densities that integrate to the total number of electrons, with the minimizing density being v-representable—corresponding to some external potential. These results shift the intractable many-body wavefunction problem to a functional minimization over three-dimensional densities, drastically reducing complexity while preserving exactness in principle. To make DFT computationally tractable, Kohn and Sham introduced in 1965 a mapping to an auxiliary system of non-interacting electrons that reproduce the interacting density \rho(\mathbf{r}). This leads to the Kohn-Sham equations, a set of single-particle orbital equations: -\frac{\nabla^2}{2} \phi_i(\mathbf{r}) + v_s(\mathbf{r}) \phi_i(\mathbf{r}) = \epsilon_i \phi_i(\mathbf{r}), solved self-consistently, where the effective potential is v_s(\mathbf{r}) = v_{\text{ext}}(\mathbf{r}) + v_H[\rho](\mathbf{r}) + v_{\text{xc}}[\rho](\mathbf{r}). Here, v_H[\rho](\mathbf{r}) = \int \frac{\rho(\mathbf{r}')}{|\mathbf{r} - \mathbf{r}'|} d\mathbf{r}' is the classical Hartree (Coulomb) potential from the density, and v_{\text{xc}}[\rho](\mathbf{r}) = \frac{\delta E_{\text{xc}}[\rho]}{\delta \rho(\mathbf{r})} derives from the exchange-correlation functional. The density is constructed as \rho(\mathbf{r}) = \sum_{i=1}^N |\phi_i(\mathbf{r})|^2, with the sum over the N occupied Kohn-Sham orbitals, yielding the total energy E = T_s[\rho] + \int v_{\text{ext}}(\mathbf{r}) \rho(\mathbf{r}) d\mathbf{r} + E_H[\rho] + E_{\text{xc}}[\rho], where T_s[\rho] is the non-interacting kinetic energy. This framework, akin to mean-field theories like Hartree but augmented by correlation effects, transforms the many-body problem into a series of one-body problems. The exchange-correlation functional E_{\text{xc}}[\rho] encapsulates all many-body quantum effects beyond mean-field electrostatics, but its exact form is unknown, necessitating approximations. The local density approximation (LDA) treats E_{\text{xc}}[\rho] by assuming the system locally resembles a homogeneous electron gas, yielding E_{\text{xc}}^{\text{LDA}}[\rho] = \int \epsilon_{\text{xc}}(\rho(\mathbf{r})) \rho(\mathbf{r}) \, d\mathbf{r}, where \epsilon_{\text{xc}}(\rho) is the per-particle exchange-correlation energy of the uniform gas, parametrized from accurate calculations. This simple yet effective approximation captures exchange exactly in the high-density limit and correlation via empirical fits, enabling practical implementations. Kohn-Sham DFT has achieved widespread success in many-body physics, routinely applied to electronic structure calculations for systems comprising hundreds of atoms, such as molecules, solids, and nanostructures, due to its favorable scaling with system size compared to wavefunction methods. Despite this, DFT exhibits limitations in strongly correlated regimes, where approximate functionals like LDA fail to adequately describe electron localization and Mott insulators, often severely underestimating band gaps in semiconductors and transition-metal compounds by factors of two or more. As of 2025, advancements include machine learning-enhanced DFT, where neural networks correct errors in approximate functionals, achieving near-chemical accuracy (e.g., reducing energy errors to ~1 kcal/mol) for diverse datasets, and new functionals like Skala XC for improved predictions in complex materials.

Applications

Condensed Matter Systems

In condensed matter physics, the many-body problem is central to understanding the collective behavior of electrons in solids, where interactions lead to emergent phenomena such as phase transitions, magnetism, and superconductivity. These systems, often extended lattices or continua of interacting fermions, exhibit properties that cannot be captured by single-particle approximations due to strong correlations and quantum statistics. Seminal models simplify the complexity while capturing essential physics, enabling predictions of macroscopic behaviors like electrical conductivity and magnetic ordering from microscopic Hamiltonians. The uniform electron gas, modeled as , provides a cornerstone for describing conduction electrons in metals, treating them as a sea of interacting particles embedded in a uniform positive background to screen long-range Coulomb repulsion. This idealized system highlights the competition between kinetic energy and electron-electron interactions, parameterized by the dimensionless density r_s = (3/4\pi n)^{1/3}/a_B, where n is the electron density and a_B the . At high densities (low r_s), the state dominates, with quasiparticles behaving like weakly interacting fermions, as developed in treatments. However, at low densities (high r_s), repulsive interactions prevail, causing electrons to localize into a lattice structure known as the , predicted in 1934, where the classical Coulomb energy outweighs quantum delocalization. Magnetic ordering in solids arises from exchange interactions between localized spins, paradigmatically captured by the Heisenberg model with Hamiltonian H = -J \sum_{\langle i,j \rangle} \mathbf{S}_i \cdot \mathbf{S}_j, where J sets the interaction strength and the sum runs over nearest neighbors. For ferromagnetic materials (J < 0), spins align parallel below the , leading to spontaneous magnetization as a collective many-body effect. In the mean-field approximation, which replaces fluctuating interactions with an average field, the Curie temperature is T_c = 2z |J| S(S+1) / 3 k_B, with z the lattice coordination number and S the spin quantum number; this yields reasonable estimates for transition temperatures in simple magnets, though it overestimates T_c by neglecting fluctuations. Antiferromagnetic cases (J > 0) exhibit staggered ordering, influencing properties like in insulators. Superconductivity exemplifies many-body pairing, where conventional mechanisms involve electron-phonon attraction overcoming Coulomb repulsion. The Bardeen-Cooper-Schrieffer (BCS) theory of 1957 explains this through the formation of Cooper pairs, bound states of time-reversed electrons near the Fermi surface, resulting in a superconducting energy gap \Delta \sim \hbar \omega_D \exp(-1/N(0)V), with \omega_D the Debye frequency, N(0) the density of states at the Fermi energy, and V the effective pairing interaction; this weak-coupling limit predicts exponential temperature dependence of the critical field and specific heat jump at T_c. For materials with stronger electron-phonon coupling (\lambda > 0.3), Eliashberg theory extends BCS by solving self-consistent Dyson equations that incorporate phonon frequency dependence and retardation, yielding larger gaps and higher T_c values, as verified in lead and niobium. These approaches underscore phase transitions from normal to superconducting states driven by collective pairing instability. Strong correlations can drive metal-insulator transitions, as in the , which balances kinetic hopping t and on-site repulsion U via H = -t \sum_{\langle i,j \rangle, \sigma} (c_{i\sigma}^\dagger c_{j\sigma} + \text{h.c.}) + U \sum_i n_{i\uparrow} n_{i\downarrow}, where c^\dagger, c are fermionic operators and n number operators. At half-filling (one per site), increasing U/t from weak values promotes a metallic state with itinerant carriers, but around U/t \sim 1, the Mott transition occurs: double occupancy is suppressed, opening a charge gap and yielding an antiferromagnetic insulator due to . This quantum , first modeled in , captures correlation-induced localization in transition metal oxides and cuprates, highlighting how bandwidth competition with repulsion fosters insulating behavior despite partially filled bands.

Nuclear and Particle Physics

In , the many-body problem manifests prominently in the description of atomic nuclei, where strong nuclear forces govern the interactions among protons and neutrons. The addresses this by treating nucleons as independent particles moving in a mean-field potential generated by all others, augmented by residual interactions that account for correlations beyond the mean field. This approach, developed in the mid-20th century, successfully explains and nuclear spectra by filling discrete energy levels analogous to atomic electron shells. Residual interactions, such as tensor and spin-orbit components from realistic nucleon-nucleon potentials, introduce two-body and higher-order effects that refine binding energies and excitation levels. Pairing correlations in the shell model further capture collective phenomena, particularly the preference for even numbers in stable nuclei. Modeled via a BCS-like adapted for nuclei, pairing treats like-nucleon pairs (protons or neutrons) as correlated Cooper pairs near the , leading to a pairing that manifests in even-odd mass differences of approximately 1-2 MeV. This arises from the attractive component of the in the spin-singlet channel, enhancing binding in even-even nuclei relative to odd-mass ones and explaining phenomena like ground-state pairing . Correlation effects in these fermionic systems amplify beyond mean-field predictions, necessitating inclusion of multi-particle excitations for quantitative accuracy. Ab initio methods provide parameter-free solutions to the nuclear many-body Hamiltonian for light nuclei, bypassing phenomenological adjustments. The no-core shell model (NCSM) diagonalizes the full intrinsic Hamiltonian \hat{H} = \sum_{i<j} V_{ij} + \sum_{i<j<k} V_{ijk} in a complete harmonic oscillator basis, using realistic two- and three-nucleon interactions derived from chiral effective field theory. This approach yields converged ground- and excited-state properties for mass number A \leq 16, such as binding energies and electromagnetic transitions in nuclei like ^{12}\mathrm{C} and ^{16}\mathrm{O}, with uncertainties below 1% for key observables when basis spaces reach N_{\max} \approx 10. For heavier systems, computational scaling limits exact solutions, prompting importance-truncated variants. For the lightest nuclei, exact few-body techniques handle the strong three-body correlations inherent to nuclear forces. The tritium (^3\mathrm{H}) binding energy of 8.48 MeV is computed precisely using Faddeev equations, which decompose the three-nucleon wave function into components satisfying coupled integral equations and incorporate two- and three-body potentials without approximation. These equations fully resolve the bound-state problem for A = 3, capturing the role of three-nucleon forces in achieving the experimental binding. However, for A > 3, such as ^4\mathrm{He}, exact Faddeev-Yakubovsky extensions become intractable due to increasing cluster , requiring hyperspherical or other approximations to manage the many-body complexity. In , the many-body problem extends to high-density regimes under (QCD), exemplified by the quark-gluon plasma (QGP) formed in relativistic heavy-ion collisions. This state consists of deconfined and behaving as a strongly interacting fluid at temperatures above the deconfinement transition, estimated at approximately 156-157 MeV (as of 2024) from simulations. discretizes to compute the QCD partition function non-perturbatively, revealing a crossover rather than first-order transition near this temperature, with thermodynamic quantities like pressure and energy density showing rapid increases due to gluon liberation. These calculations, performed on quark masses tuned to physical values, validate QGP properties such as shear viscosity over entropy ratio close to the AdS/CFT bound, informing experimental observations at facilities like RHIC and LHC.

References

  1. [1]
    [PDF] Case Studies in Many-Body Physics - ODU Digital Commons
    The many-body problem refers to any physical problem made of more than two interacting particles. With increasing number of particles in a system, their cou.
  2. [2]
    [PDF] N-BODY PROBLEMS Math118, O. Knill
    In this section, we look at general facts about the n-body problem like existence of solutions and the nature of singularities like Zeipel theorem which ...
  3. [3]
    [PDF] Quantum Many-Body Physics in a Nutshell - Chapter 1
    This book covers quantum many-body physics, including phenomena like superfluidity, superconductivity, and tools like Feynman diagrams, and systems like ...
  4. [4]
    Stochastic representation of many-body quantum states - PMC
    Jun 16, 2023 · The quantum many-body problem is ultimately a curse of dimensionality: the state of a system with many particles is determined by a function ...<|control11|><|separator|>
  5. [5]
    [PDF] N-body Problem: Analytical and Numerical Approaches - UNOOSA
    We present an over view of the Hamiltonian of the N-Body problem with some special cases (two- and three-body problems) in view of classical mechanics and ...
  6. [6]
    Many-body problem - Oxford Reference
    The problem that it is very difficult to obtain exact solutions to systems involving interactions between more than two bodies – using either classical ...Missing: authoritative | Show results with:authoritative
  7. [7]
    [PDF] Quantum Theory of Many Particle Systems
    Mar 28, 2023 · 1The definition has an extra factor (-1)n, to be consistent with Fetter-Walecka's definition. See FW Eq. (23.6). 16. Page 25. Proof. G (α0; α.
  8. [8]
    Les méthodes nouvelles de la mécanique céleste - Internet Archive
    May 6, 2010 · Les méthodes nouvelles de la mécanique céleste ; Publication date: 1892 ; Topics: Celestial mechanics ; Publisher: Paris : Gauthier-Villars et fils.Missing: body integrability
  9. [9]
    [PDF] Poincaré and the Three-Body Problem
    The Three-Body Problem has been a recurrent theme of Poincaré's thought. Having understood very early the need for a qualitative study of “non- integrable” ...
  10. [10]
    Elementary Principles in Statistical Mechanics
    Josiah Willard Gibbs. Publisher: Cambridge University Press. Online publication date: August 2011. Print publication year: 2010. First published in: 1902.
  11. [11]
    Werner Heisenberg: the Columbus of quantum mechanics
    Dec 4, 2001 · This year marks the centenary of the birth of Werner Heisenberg, pioneer of quantum mechanics and theoretical high-energy physics.
  12. [12]
    Quantum mechanics of many-electron systems - Journals
    To get agreement with experiment it was found necessary to introduce the spin of the electron, giving a doubling in the number of orbits of an electron in an ...
  13. [13]
    [PDF] Hans Bethe: The Nuclear Many Body Problem - arXiv
    May 19, 2016 · Now atomic physics was considered to be very different from nuclear physics, the atomic many body problem being that of explaining the makeup of ...
  14. [14]
    When condensed-matter physics became king
    Jan 1, 2019 · “Condensed matter” called to mind the notoriously difficult quantum many-body calculations more than “solid state,” and trends during the 1960s ...Missing: post- | Show results with:post-
  15. [15]
    Space-Time Approach to Non-Relativistic Quantum Mechanics
    The contribution from a single path is postulated to be an exponential whose (imaginary) phase is the classical action (in units of ) for the path in question.
  16. [16]
    [PDF] Classical Mechanics - Rutgers Physics
    Nov 17, 2010 · Classical mechanics, narrowly defined, is the investigation of the motion of systems of particles in Euclidean three-dimensional space, ...
  17. [17]
  18. [18]
    [PDF] Liouville–Arnold theorem for homogeneous and contact Hamiltonian ...
    A Hamiltonian system is completely integrable (in the sense of Liouville) if there exist as many independent integrals of motion in involution as the dimension ...
  19. [19]
    [PDF] Liouville-Arnold theorems
    Liouville, proved the first ”integrability by quadratures” theorem which in modern terms [33] can be formulated as fol- lows. Theorem 0.1. Let M2n ≃ T∗(Rn) be a ...
  20. [20]
    Liouville Integrability
    The Liouville theorem in classical mechanics states the conditions under which the equations of motion of a dynamical system can always be solved by means of a ...
  21. [21]
    [PDF] Integrability and non integrability of some n body problems - arXiv
    Sep 28, 2015 · . Already since Poincare and Bruns [1, 2], it is known that the n-body problem is for n ≥ 3 not integrable in general. Bruns in [1] proved ...
  22. [22]
    A new proof of Poincaré's result on the restricted three-body problem
    Nov 22, 2021 · Abstract:The problem of nonintegrability of the circular restricted three-body problem is very classical and important in dynamical systems.
  23. [23]
    Lyapunov Spectrum Scaling for Classical Many-Body Dynamics ...
    Long-range limits result in a single parameter scaling of the Lyapunov spectrum, with the inverse largest Lyapunov exponent being the only diverging control ...
  24. [24]
  25. [25]
    The quantum theory of the emission and absorption of radiation
    The quantum theory of the emission and absorption of radiation. Paul Adrien Maurice Dirac.
  26. [26]
    [PDF] P. Jordan, E. Wigner. Über das Paulischen Äquivalenzverbot
    (5). Über das Paulische Äquivalenzverbot. Von P. Jordan und E. Wigner in Göttingen. (Eingegangen am 26. Januar 1928.) Die Arbeit enthält eine Fortsetzung der ...Missing: Äquivalenzproblem | Show results with:Äquivalenzproblem
  27. [27]
    [PDF] E{i
    2. ha Myne 15, 622 (1930). Konfigurationsraum und zweite Quantelung ¹). Von V. Fock in Leningrad. (Eingegangen am 10. März 1932.) Es wird der Zusammenhang ...
  28. [28]
    [PDF] Chapter 2 Second Quantisation - TCM
    Second quantization is a method for formulating many-body quantum systems, providing an efficient language and a way to condense the representation of ...
  29. [29]
    [PDF] SECOND QUANTIZATION Lecture notes with course Quantum Theory
    Second quantization is the standard formulation of quantum many-particle theory, important for Quantum Field and Condensed Matter Theory.
  30. [30]
    [PDF] Chapter 4 Introduction to many-body quantum mechanics
    For bosons the general many-body wave function thus needs to be symmetric under ... For fermions the wave function has to be antisymmetric under exchange ...
  31. [31]
    [PDF] Modelling equilibration of local many-body quantum systems by ...
    May 28, 2020 · For a many-body system with large Hilbert space dimension, the dimension of the equilibrium subspace is typically much larger than the dimension.
  32. [32]
    Chemistry beyond the scale of exact diagonalization on a quantum ...
    Jun 18, 2025 · This limits exact diagonalization in the full configuration interaction (FCI) to system sizes close to 22 electrons in 22 orbitals (22e,22o) (1) ...
  33. [33]
    Easing the Monte Carlo sign problem | Science Advances
    Aug 14, 2020 · INTRODUCTION. Quantum Monte Carlo (QMC) techniques are central to our understanding of the equilibrium physics of many-body quantum systems.
  34. [34]
    [PDF] The Phase Space Elementary Cell in Classical and ... - arXiv
    In classical statistical mechanics, the statement (∆Ω)N = h3N is justified considering the product of an N independent quantum one-particle system or from the ...Missing: body | Show results with:body
  35. [35]
    [quant-ph/0406180] The Complexity of the Local Hamiltonian Problem
    Jun 24, 2004 · It was known that the problem is QMA-complete for any k <= 3. On the other hand 1-local Hamiltonian is in P, and hence not believed to be QMA- ...Missing: finding ground
  36. [36]
    Correlation Problem in Many‐Electron Quantum Mechanics I ...
    This chapter contains sections titled: Introduction Formulation of the Correlation Problem Methods for Treating Electronic Correlation Recent Developments; ...
  37. [37]
    Coulomb Hole About a Light Test Charge in an Electron Gas at ...
    Coulomb Hole About a Light Test Charge in an Electron Gas at Metallic Densities | Phys. Rev. Lett.
  38. [38]
    The Configuration Interaction Method: Advances in Highly ...
    We review the fundamentals of the configuration interaction method and discuss various determinant-based CI algorithms.Missing: strong seminal paper
  39. [39]
    Pair-correlation function of the electron gas with long-range ...
    Jan 15, 1998 · We present analytical and numerical results for the pair-correlation function $g(r=0)$ of the electron gas with long-range Coulomb ...
  40. [40]
    Strong correlation in incremental full configuration interaction
    Jun 12, 2017 · Incremental Full Configuration Interaction (iFCI) reaches high accuracy electronic energies via a many-body expansion of the correlation energy ...
  41. [41]
    A mathematical and computational review of Hartree-Fock SCF ...
    May 2, 2007 · We present here a review of the fundamental topics of Hartree-Fock theory in Quantum Chemistry. From the molecular Hamiltonian, using and discussing the Born- ...Missing: seminal | Show results with:seminal
  42. [42]
    The Wave Mechanics of an Atom with a Non-Coulomb Central Field ...
    Oct 24, 2008 · The paper is concerned with the practical determination of the characteristic values and functions of the wave equation of Schrodinger for a non-Coulomb ...Missing: Douglas | Show results with:Douglas
  43. [43]
    [PDF] 43. A STATISTICAL METHOD FOR THE DETERMINATION OF ...
    Jun 8, 2005 · The purpose of this work is to show some results about the distribution of electrons in a heavy atom which can be obtained dealing with ...
  44. [44]
    Thomas, L. (1927) The Calculation of Atomic Fields. Mathematical ...
    Thomas-Fermi theory is an approximate method, which is widely used to describe the properties of matter at various hierarchical levels.Missing: original | Show results with:original
  45. [45]
    [PDF] many-particle schrödinger equation - MSU chemistry
    MANY-BODY PERTURBATION THEORY AND ITS. DIAGRAMMATIC ... Rayleigh-Schrödinger perturbation theory, wave, reaction, and reduced resolvent operators.
  46. [46]
  47. [47]
    Quantum Monte Carlo simulations of solids | Rev. Mod. Phys.
    Jan 5, 2001 · This article describes the variational and fixed-node diffusion quantum Monte Carlo methods and how they may be used to calculate the properties of many- ...
  48. [48]
  49. [49]
  50. [50]
  51. [51]
  52. [52]
    Inhomogeneous Electron Gas | Phys. Rev.
    This paper deals with the ground state of an interacting electron gas in an external potential, and proves a universal functional of the density.
  53. [53]
    Self-Consistent Equations Including Exchange and Correlation Effects
    Self-Consistent Equations Including Exchange and Correlation Effects. W. Kohn and L. J. Sham ... 140, A1133 – Published 15 November, 1965. DOI: https://doi ...
  54. [54]
    Density functional theory: Its origins, rise to prominence, and future
    Aug 25, 2015 · Density functional theory has been spectacularly successful in physics, chemistry, and related fields, and it keeps finding new applications ...
  55. [55]
    [PDF] MEAN-FIELD APPROACH TO MAGNETISM
    In the vicinity of the critical temperature, however the experimentally obtained critical exponent for the susceptibility of ferromagnetic materials is γ ≈1.3.Missing: 2zJS²/ 3k_B
  56. [56]
    Theory of Superconductivity | Phys. Rev.
    A theory of superconductivity is presented, based on the fact that the interaction between electrons resulting from virtual exchange of phonons is attractive.
  57. [57]
    Electron correlations in narrow energy bands - Journals
    A simple, approximate model for the interaction of electrons in narrow energy bands is introduced. The results of applying the Hartree-Fock approximation to ...
  58. [58]
    The shell model as a unified view of nuclear structure
    Jun 16, 2005 · Our aim is to give the reader an overall view of the shell model as a subdiscipline of the many-body problem. ... nuclear many-body problem for ...Missing: seminal | Show results with:seminal
  59. [59]
    Odd-even mass difference and isospin dependent pairing interaction
    Aug 31, 2009 · The neutron and proton odd-even mass differences are studied with Hartree-Fock + BCS ( H F + B C S ) calculations with Skyrme interactions ...Missing: like | Show results with:like
  60. [60]
    Nuclear pairing models - ScienceDirect.com
    Jan 6, 1992 · By comparing the calculated pairing gaps to experimental odd-even mass differences we determine parameter values for the proposed forms of ...
  61. [61]
    [0904.0463] Recent developments in no-core shell-model calculations
    Apr 2, 2009 · We present an overview of recent results and developments of the no-core shell model (NCSM), an ab initio approach to the nuclear many-body problem for light ...
  62. [62]
    The ab initio No-Core Shell Model and Light Nuclei
    Oct 16, 2010 · The ab initio no-core shell model (NCSM) is a well-established theoretical framework aimed at an exact description of nuclear structure ...
  63. [63]
    Relativistic effects in the three-nucleon bound-state problem
    Jan 1, 1981 · We calculate the binding energy of tritium using two relativistic versions of the Faddeev equations with separable potentials, ...
  64. [64]
    Recent lattice QCD results and phase diagram of strongly interacting ...
    It is now well established that the QCD transition is just a rapid crossover at a transition temperature T c ∼ 155 MeV [10], [11]. Its chiral aspects manifest ...
  65. [65]
    THE SEARCH FOR THE QUARK-GLUON PLASMA - Annual Reviews
    ... temperature lies in the range 150 ± 10. MeV at vanishing net quark density. According to the standard cosmological model (8), the temperature of the cosmic ...