Fact-checked by Grok 2 weeks ago

Indistinguishable particles

In , indistinguishable particles, also referred to as identical particles, are entities that cannot be differentiated from one another by any physical measurement or labeling, even in principle, due to their identical intrinsic such as , charge, and . This fundamental property necessitates that the quantum states of such particles be described using wave functions that are either fully symmetric or fully antisymmetric under particle exchange, as dictated by the symmetrization postulate. Particles are classified into two categories based on their intrinsic spin: bosons, which have integer spin values (0, 1, 2, etc.) and form symmetric states under permutation, and fermions, which have half-integer spin values (1/2, 3/2, etc.) and form antisymmetric states. For bosons, this symmetry allows multiple particles to occupy the same quantum state, leading to phenomena like Bose-Einstein condensation, while for fermions, antisymmetry enforces the Pauli exclusion principle, which prohibits more than one particle from occupying the same state, a key factor in the structure of atoms and matter. The permutation operator for exchanging two particles yields an eigenvalue of +1 for bosons and -1 for fermions, ensuring that physical observables remain invariant under such exchanges. This indistinguishability underpins the of , with bosons following Bose-Einstein statistics and fermions adhering to Fermi-Dirac statistics, which are essential for describing collective behaviors in gases, solids, and exotic states of matter. In multi-particle systems, the is constructed from symmetrized or antisymmetrized products of single-particle states, resolving degeneracies that would arise in classical distinguishable treatments and enabling accurate predictions of energy levels and interactions.

Classical Perspective

Distinguishing particles in classical mechanics

In , particles are treated as distinguishable entities, each uniquely identified by their positions, velocities, and resulting trajectories through space-time. This assumption allows for the labeling of individual particles, such that permuting their labels does not alter the physical description of the system, as each particle's path can be continuously tracked in principle. Such distinguishability underpins the formulation of classical dynamics, where the state of a system is specified by the complete set of coordinates and momenta for all particles. Representative examples illustrate this classical perspective. In a game of billiards, each maintains a distinct through its unique and at any instant, enabling of collisions and paths without . Similarly, in the solar system are distinguished by their orbital trajectories, positions, and velocities, allowing astronomers to model their motions as separate entities governed by Newtonian . This tracking preserves the particles' individuality, even in complex interactions. Historically, Ludwig Boltzmann's early work in , such as his 1868 contributions to the , relied on treating particles as distinguishable to derive distribution functions like the Maxwell-Boltzmann . This approach facilitated the connection between microscopic motions and macroscopic but introduced inconsistencies, notably the . In the classical , assuming distinguishability leads to additive entropies when mixing two samples of the same gas, resulting in a finite increase (e.g., \Delta S = 2 N k \ln 2 for two samples each with N particles and equal volumes V) despite no observable change, and yields non-extensive that violates thermodynamic scalability. The paradox highlights the limitations of classical distinguishability, later addressed in through particle indistinguishability.

Transition to quantum indistinguishability

In , particles are distinguishable by tracking their unique trajectories and interactions over time. However, introduces fundamental limitations that undermine this approach. The Heisenberg uncertainty principle, which states that the product of the uncertainties in a particle's position \Delta x and momentum \Delta p satisfies \Delta x \Delta p \geq \hbar/2, prevents precise determination of both quantities simultaneously, making it impossible to label particles by their paths without significantly disturbing the system. Complementing this, wave-particle duality implies that particles like electrons exhibit wave-like properties, with a de Broglie wavelength \lambda = h/p that causes delocalization and interference, further prohibiting the classical notion of distinct, traceable identities. These principles collectively render trajectory-based distinguishability untenable in quantum systems. A illustrative thought experiment involves sending two electrons through a Stern-Gerlach apparatus, which measures spin along a magnetic field gradient. Each electron randomly deflects to either the spin-up or spin-down detector with 50% probability, but the results are statistically identical regardless of any attempt to assign labels like "electron A" or "electron B" beforehand. Since no measurement can track or differentiate them without altering their —due to the in position during deflection—the experiment reveals that the particles produce indistinguishable outcomes, implying they lack individual identities. This highlights how quantum measurements treat such particles as fundamentally unlabeled. Identical particles in are those sharing the same intrinsic properties, including , charge, , and other quantum numbers, such that no physical interaction or observable can distinguish one from another, even in principle. For instance, all s are identical because they possess identical values for these properties, and experiments like or yield the same results irrespective of which specific electron is involved. This definition contrasts with classical particles, where subtle differences in initial conditions or external marks allow differentiation. The conceptual shift toward quantum indistinguishability emerged in early quantum developments. In his 1924 paper on the quantum theory of an ideal monatomic gas, Albert Einstein extended Satyendra Nath Bose's statistics for photons to material particles, emphasizing their indistinguishability to resolve issues like the Gibbs paradox in entropy calculations. In a follow-up 1925 paper, Einstein further developed the quantum theory of the monatomic ideal gas using Bose-Einstein statistics. This exchange symmetry framework became the formal tool for handling indistinguishability in quantum theory.

Quantum Indistinguishability

Definition and motivation

In , indistinguishable particles are identical particles—such as electrons or photons—that share all intrinsic properties like , charge, and , and cannot be differentiated by any physical measurement, even in principle. This indistinguishability arises from the symmetrization postulate, which requires the multi-particle state in the to transform in a specific way under particle label exchanges, ensuring that no depends on arbitrary labeling. Unlike classical particles, which can be tracked via trajectories, quantum particles lack definite paths, making individual identification impossible and leading to invariance as a fundamental requirement. The motivation for this treatment stems from the need to avoid unphysical predictions in multi-particle systems. If particles were treated as distinguishable, quantum calculations would overcount degenerate states by including permutations as distinct, violating that observables must be independent of labeling conventions. This ensures consistent physical predictions, such as correct statistical weights in ensembles, and resolves ambiguities in state descriptions due to degeneracy. Historically, the classical —where mixing identical ideal gases leads to non-extensive unless particles are deemed indistinguishable—served as a precursor, highlighting the inadequacy of classical distinguishability for thermodynamic consistency. A key physical implication appears in atomic systems, where indistinguishability directly influences energy levels. For electrons, which are fermions, in the , the two identical electrons require an antisymmetric total wavefunction; in the ( spin configuration), the spatial part is symmetric, leading to higher repulsion and elevated energy, while excited triplet states have antisymmetric spatial wavefunctions that increase average separation, lowering energy compared to naive distinguishable estimates. In contrast, atoms, being composite bosons, can all occupy the same in , enabling without such exclusion effects. Permutation symmetry ensures that the exchange eigenvalue is conserved because the commutes with permutation operators, with the core driver being enabling entangled, non-local descriptions of identical particles.

Exchange operator and symmetry

In , the formalizes the consequences of particle indistinguishability by implementing permutations of particle labels, ensuring that physical predictions remain unchanged under such relabelings. For a system of two indistinguishable particles, the exchange operator P_{12} is defined to swap the spatial coordinates and internal degrees of freedom, such as spin, between the particles. It acts on the two-particle wavefunction as P_{12} \psi(\mathbf{x}_1, \mathbf{x}_2) = \psi(\mathbf{x}_2, \mathbf{x}_1), where \mathbf{x}_i = (\mathbf{r}_i, s_i) denotes the position \mathbf{r}_i and spin s_i of the i-th particle. This operator satisfies P_{12}^2 = I, the identity, and its eigenvalues are \pm 1. States with eigenvalue +1 are symmetric under exchange, characteristic of bosons, while those with eigenvalue -1 are antisymmetric, characteristic of fermions. For identical particles, the total wavefunction must be an eigenstate of the P_{12}, with the eigenvalue fixed by the : +1 for bosons and -1 for fermions. This ensures that the wavefunction respects the underlying required for indistinguishable particles. For systems of N > 2 particles, the wavefunction must similarly be an eigenstate of all pairwise exchange operators P_{ij} (for i < j), corresponding to a representation of the full symmetric group S_N. The requirement of exchange symmetry follows from a fundamental postulate: all physical observables, including the Hamiltonian H, must commute with every permutation operator in S_N, i.e., [H, P] = 0 for any P \in S_N. Consequently, the energy eigenstates of H are also eigenstates of these permutations, enforcing symmetric or antisymmetric behavior depending on the particle type. This commutativity reflects the physical invariance under particle relabeling for truly identical particles.

Wavefunction Formalism

Symmetric and antisymmetric wavefunctions

In the formalism of quantum mechanics for indistinguishable particles, the multi-particle wavefunction must exhibit definite symmetry under the exchange of particle labels to account for their indistinguishability: it remains unchanged (symmetric) for and acquires a minus sign (antisymmetric) for . For two occupying distinct single-particle states \psi_a(\mathbf{r}) and \psi_b(\mathbf{r}) that are orthonormal (\langle \psi_a | \psi_b \rangle = 0), the symmetric spatial wavefunction is constructed as the normalized symmetrized product: \psi_s(\mathbf{r}_1, \mathbf{r}_2) = \frac{1}{\sqrt{2}} \left[ \psi_a(\mathbf{r}_1) \psi_b(\mathbf{r}_2) + \psi_a(\mathbf{r}_2) \psi_b(\mathbf{r}_1) \right]. This ensures \int |\psi_s(\mathbf{r}_1, \mathbf{r}_2)|^2 \, d\mathbf{r}_1 \, d\mathbf{r}_2 = 1. If both bosons occupy the same state \psi_a, the wavefunction simplifies to \psi_a(\mathbf{r}_1) \psi_a(\mathbf{r}_2), which is inherently symmetric and normalized given the single-particle normalization. For two fermions in orthonormal states \psi_a and \psi_b, the antisymmetric spatial wavefunction is: \psi_{as}(\mathbf{r}_1, \mathbf{r}_2) = \frac{1}{\sqrt{2}} \left[ \psi_a(\mathbf{r}_1) \psi_b(\mathbf{r}_2) - \psi_a(\mathbf{r}_2) \psi_b(\mathbf{r}_1) \right], which also satisfies normalization under the orthogonality condition and vanishes identically if \psi_a = \psi_b, preventing multiple from occupying the same state. When share identical spatial orbitals but differ in spin projections, the full antisymmetric wavefunction is expressed via a to enforce the required exchange symmetry. The normalization factor \frac{1}{\sqrt{2}} holds specifically for orthonormal single-particle states; for non-orthogonal states with overlap S = |\langle \psi_a | \psi_b \rangle|^2 < 1, it generalizes to \frac{1}{\sqrt{2(1 \pm S)}}, where the plus sign applies to symmetric wavefunctions and the minus to antisymmetric ones, preserving \int |\psi|^2 \, d\mathbf{r}_1 \, d\mathbf{r}_2 = 1. For spin-1/2 particles like , the total wavefunction is the product of spatial and spin components, and must be overall antisymmetric. The spin part is antisymmetric in the singlet state (total spin S=0) and symmetric in the triplet state (S=1), so the spatial part must be symmetric for the singlet and antisymmetric for the triplet to achieve total antisymmetry.

Representation for two particles

For two indistinguishable particles, the wavefunction is constructed by applying the appropriate symmetrization or antisymmetrization to the product of single-particle states, ensuring overall symmetry for bosons or antisymmetry for fermions. This representation captures the essential quantum effects of indistinguishability in simple systems. Consider two identical bosons confined to a one-dimensional infinite square well potential of length L. The single-particle eigenfunctions are \phi_n(x) = \sqrt{2/L} \sin(n \pi x / L), with energies E_n = (n^2 \pi^2 \hbar^2)/(2 m L^2). For the ground state, both bosons can occupy the lowest orbital (n=1), yielding the symmetric two-particle wavefunction \psi(x_1, x_2) = \phi_1(x_1) \phi_1(x_2). The probability density |\psi(x_1, x_2)|^2 = |\phi_1(x_1)|^2 |\phi_1(x_2)|^2 is symmetric under particle exchange, allowing enhanced occupation of the same state and bunching effects observable in interference experiments. In contrast, for two identical fermions such as electrons in the same potential well, the total wavefunction (including spin) must be antisymmetric to satisfy the . The ground state configuration places one fermion in the n=1 spatial orbital with spin up and the other in the n=1 spatial orbital with spin down, forming an antisymmetric spin singlet state \chi_\text{singlet} = (\alpha(1)\beta(2) - \beta(1)\alpha(2))/\sqrt{2}, where \alpha and \beta denote spin up and down. The spatial part remains symmetric, \phi_1(x_1) \phi_1(x_2), so the overall wavefunction \psi(x_1, x_2, \sigma_1, \sigma_2) = \phi_1(x_1) \phi_1(x_2) \chi_\text{singlet} is antisymmetric. If spins were parallel (triplet, symmetric spin), the spatial wavefunction would need to be antisymmetric, such as [\phi_1(x_1) \phi_2(x_2) - \phi_1(x_2) \phi_2(x_1)] / \sqrt{2}, raising the energy. This enforces single occupancy per spin-orbital, preventing both fermions from sharing the ground spatial state with identical spins. The indistinguishability also introduces an exchange energy shift in the expectation value of the Hamiltonian for interacting particles. For a two-particle system with Hamiltonian H = h(1) + h(2) + V(| \mathbf{r}_1 - \mathbf{r}_2 |), where h is the single-particle operator, the energy is \langle H \rangle = E_0 \pm J. Here, E_0 includes the sum of single-particle energies and the direct Coulomb integral, while J = \int \phi_a^*(\mathbf{r}_1) \phi_b^*(\mathbf{r}_2) V(|\mathbf{r}_1 - \mathbf{r}_2|) \phi_a(\mathbf{r}_2) \phi_b(\mathbf{r}_1) d\mathbf{r}_1 d\mathbf{r}_2 is the exchange integral, with the sign + for symmetric (boson or singlet fermion) states and - for antisymmetric ones. This interference term lowers the energy for antisymmetric spatial wavefunctions, stabilizing triplet states relative to singlets in some systems. A prototypical illustration is the ground state of the helium atom, where both electrons occupy the 1s hydrogenic orbital \phi_{1s}(\mathbf{r}) = (1/\sqrt{\pi}) (Z/a_0)^{3/2} e^{-Z r / a_0} (with effective Z \approx 1.7). The spatial wavefunction is symmetric, \psi_\text{spatial}(\mathbf{r}_1, \mathbf{r}_2) = \phi_{1s}(\mathbf{r}_1) \phi_{1s}(\mathbf{r}_2), paired with an antisymmetric spin singlet \chi_\text{singlet}, yielding total energy approximately -79 eV. This configuration minimizes energy by maximizing spatial overlap while satisfying antisymmetry, with the exchange integral contributing to the correlation effects beyond independent-particle approximations.

Extension to N particles

The extension to systems of N indistinguishable particles requires the total wavefunction to belong to specific irreducible representations of the permutation group S_N, which consists of all N! possible exchanges of particle labels. For bosons, the wavefunction must transform under the fully symmetric representation of S_N, ensuring invariance under any permutation, while for fermions, it must transform under the fully antisymmetric representation, acquiring a phase of (-1)^p where p is the parity of the permutation. This generalization enforces the symmetrization postulate for arbitrary N, as originally proposed by Dirac for identical particles in quantum mechanics. For N bosons occupying single-particle orbitals \phi_i(\mathbf{r}), the wavefunction is given by the symmetrized product, explicitly constructed as the permanent of the matrix whose elements are these orbitals evaluated at the particle coordinates: \Psi_B(\mathbf{r}_1, \dots, \mathbf{r}_N) = \frac{1}{\sqrt{N!}} \sum_{P \in S_N} \prod_{k=1}^N \phi_{P(k)}(\mathbf{r}_k), where the sum runs over all permutations P in S_N. This form, analogous to the two-particle symmetric combination but extended to full symmetrization, allows multiple bosons to occupy the same state. In contrast, for N fermions, the wavefunction is the antisymmetrized product, represented by the Slater determinant to ensure antisymmetry and the Pauli exclusion principle, preventing more than one fermion per state: \Psi_F(\mathbf{r}_1, \dots, \mathbf{r}_N) = \frac{1}{\sqrt{N!}} \det \begin{vmatrix} \phi_1(\mathbf{r}_1) & \phi_1(\mathbf{r}_2) & \cdots & \phi_1(\mathbf{r}_N) \\ \phi_2(\mathbf{r}_1) & \phi_2(\mathbf{r}_2) & \cdots & \phi_2(\mathbf{r}_N) \\ \vdots & \vdots & \ddots & \vdots \\ \phi_N(\mathbf{r}_1) & \phi_N(\mathbf{r}_2) & \cdots & \phi_N(\mathbf{r}_N) \end{vmatrix}. This determinant form, introduced by Slater as a practical way to enforce antisymmetry in multi-electron systems, vanishes if any two particles occupy identical orbitals. Exact computation of these N-particle wavefunctions scales exponentially with N due to the N! terms in the sums or determinants, rendering full solutions intractable for large systems. This computational challenge motivates approximations such as the , which assumes a single Slater determinant for fermions (or permanent for bosons) and variationally optimizes the orbitals to approximate the ground state energy.

Particle Classifications

Bosons: Properties and examples

Bosons are elementary or composite particles characterized by integer values of intrinsic angular momentum, or spin (0, 1, 2, ...), which dictates their behavior under quantum mechanical exchange operations. According to the spin-statistics theorem, the multi-particle wave function of identical bosons remains symmetric upon interchanging any two particles, allowing for constructive interference in quantum states. This symmetry arises fundamentally from the relativistic quantum field theory framework and ensures that bosonic statistics apply to all such particles. The theorem was first rigorously established in 1940, linking spin to statistical behavior without exceptions in local quantum field theories. The foundational insight into bosonic statistics emerged from Satyendra Nath Bose's 1924 derivation of , where he treated photons as indistinguishable entities rather than classical waves or distinguishable particles. Bose's approach counted microstates by assuming photons of the same energy are identical, leading to a novel distribution function that matched experimental spectra without invoking ad hoc assumptions. Albert Einstein extended this in 1924–1925 to non-relativistic ideal gases of massive particles, predicting and the possibility of condensation phenomena. This work laid the groundwork for understanding bosons beyond photons, influencing quantum statistical mechanics profoundly. Prominent examples of bosons include photons, which carry spin 1 and mediate the electromagnetic force as fundamental gauge bosons in the . Composite bosons, such as helium-4 atoms, exhibit total spin 0 due to the pairing of two protons, two neutrons, and two electrons, enabling superfluidity at low temperatures. In superconductors, Cooper pairs—bound states of two electrons with opposite momenta and spins—behave as effective spin-0 bosons, allowing collective motion without resistance as described in the . A defining property of bosons is the absence of occupancy restrictions in quantum states, permitting multiple particles to share the same single-particle state, in contrast to . This leads to , a phase transition where, below a critical temperature, a macroscopic fraction of bosons occupies the system's ground state, manifesting as coherent quantum behavior observable in dilute atomic gases and superfluids. The first experimental realization of in a dilute vapor of confirmed this macroscopic occupation, validating predictions from .

Fermions: Properties and Pauli exclusion

Fermions are subatomic particles characterized by half-integer spin values, such as \frac{1}{2}, \frac{3}{2}, and so on, which leads them to obey and possess antisymmetric total wavefunctions under particle exchange. This antisymmetric property arises from the , which connects the intrinsic spin of a particle to the symmetry requirements of its multiparticle wavefunction in relativistic . As a consequence, the wavefunction changes sign upon interchanging any two identical fermions, ensuring that the overall description remains consistent with quantum mechanical principles. The Pauli exclusion principle states that no two identical fermions can occupy the same quantum state simultaneously, meaning they cannot share the same set of quantum numbers, such as the principal quantum number n, orbital angular momentum l, magnetic quantum number m_l, and spin projection m_s. This principle, originally formulated to explain the structure of atomic electron shells and spectral complexities, prohibits configurations where fermions would violate antisymmetry. In practice, it manifests in systems like atoms, where electrons fill orbitals according to these rules, leading to the discrete shell structure observed in the periodic table of elements. Prominent examples of fermions include leptons like electrons and baryons such as protons and neutrons, all of which exhibit half-integer spin and adhere to the exclusion principle in their interactions and bound states. For instance, in atomic physics, the exclusion principle governs electron occupancy in subshells, with a maximum of two electrons per orbital (one for each spin state), directly accounting for the chemical properties and periodicity of elements. The exclusion principle derives directly from the antisymmetry of the fermionic wavefunction, typically represented as a Slater determinant for a system of N fermions: \Psi(1, 2, \dots, N) = \frac{1}{\sqrt{N!}} \det \begin{pmatrix} \phi_1(1) & \phi_1(2) & \cdots & \phi_1(N) \\ \phi_2(1) & \phi_2(2) & \cdots & \phi_2(N) \\ \vdots & \vdots & \ddots & \vdots \\ \phi_N(1) & \phi_N(2) & \cdots & \phi_N(N) \end{pmatrix}, where \phi_i are single-particle spin-orbitals. If any two fermions occupy identical states, say \phi_i = \phi_j for i \neq j, two rows of the determinant become identical, causing the entire determinant—and thus the wavefunction—to vanish, rendering such a state impossible. This mathematical enforcement underscores the principle's role in stabilizing matter, from atomic electrons to nuclear protons and neutrons.

Multi-Particle Descriptions

Second quantization approach

Second quantization provides a powerful formalism for describing systems of indistinguishable particles, especially in many-body quantum mechanics, by reformulating the theory in terms of operators acting on an infinite-dimensional Hilbert space. Originally developed by in 1927 to quantize the electromagnetic field and treat photons as identical bosons, this approach extends naturally to both bosonic and fermionic particles, facilitating the handling of variable particle numbers without explicit reference to individual particle coordinates. Central to second quantization is the Fock space, constructed as the direct sum of Hilbert spaces for all possible particle numbers N = 0, 1, 2, \dots: \mathcal{F} = \bigoplus_{N=0}^\infty \mathcal{H}_N, where \mathcal{H}_N is the appropriately symmetrized N-particle Hilbert space. This structure, formalized by in 1932, allows states to be represented by occupation numbers in single-particle basis states, with the vacuum state |0\rangle serving as the starting point and multi-particle states generated via creation operators a^\dagger_k applied to the vacuum, where k labels orthonormal single-particle states. The dynamics are governed by creation a^\dagger_k and annihilation a_k operators, whose algebra enforces the indistinguishability. For bosons, they obey the commutation relations [a_k, a^\dagger_l] = \delta_{kl}, \quad [a_k, a_l] = [a^\dagger_k, a^\dagger_l] = 0, while for fermions, they satisfy anticommutation relations \{a_k, a^\dagger_l\} = \delta_{kl}, \quad \{a_k, a_l\} = \{a^\dagger_k, a^\dagger_l\} = 0. These relations, rooted in early developments by for bosons and extended to fermions by and in 1928, ensure that bosonic states are symmetric and fermionic states antisymmetric under particle exchange. In this framework, the Hamiltonian for a system of non-interacting indistinguishable particles takes the form H = \sum_k \epsilon_k a^\dagger_k a_k, where \epsilon_k is the single-particle energy for state k; interaction terms, such as two-body potentials, are incorporated via products like \sum_{k,l,m,n} V_{klmn} a^\dagger_k a^\dagger_l a_m a_n. This operator representation simplifies calculations for many-body systems, as seen in derivations of mean-field equations like the . A primary advantage of second quantization lies in its automatic enforcement of exchange symmetry through the operator algebra, eliminating the need for manual construction and symmetrization of multi-particle wavefunctions, which becomes impractical for large N. This shift from coordinate-based descriptions to an occupation-number basis streamlines the treatment of identical particles in quantum field theory and condensed matter applications.

Creation and annihilation operators

In the framework of second quantization, creation and annihilation operators provide a powerful method to construct states of indistinguishable particles while conserving particle number and incorporating quantum statistics. These operators act on a Fock space, enabling the description of arbitrary occupation numbers for each single-particle mode. For bosons, the annihilation operator a_{\mathbf{k}} applied to the vacuum state |0\rangle yields zero, i.e., a_{\mathbf{k}} |0\rangle = 0, reflecting the absence of particles in any mode. The corresponding number states for a single mode \mathbf{k} are constructed as |n_{\mathbf{k}}\rangle = \frac{(a_{\mathbf{k}}^\dagger)^{n_{\mathbf{k}}}}{\sqrt{n_{\mathbf{k}}!}} |0\rangle, where a_{\mathbf{k}}^\dagger is the creation operator that adds a boson to mode \mathbf{k}, and n_{\mathbf{k}} can be any non-negative integer. These operators satisfy the bosonic commutation relation [a_{\mathbf{k}}, a_{\mathbf{k}}^\dagger] = 1, ensuring symmetric wavefunctions under particle exchange. This formalism, originally developed for quantized radiation fields, extends naturally to massive bosonic particles. In contrast, for fermions, the creation and annihilation operators obey anticommutation relations \{a_{\mathbf{k}}, a_{\mathbf{k}}^\dagger\} = 1, which enforces the . Consequently, the occupation number n_{\mathbf{k}} for each mode is restricted to 0 or 1, as (a_{\mathbf{k}}^\dagger)^2 = 0, preventing multiple occupancy. The fermionic vacuum remains a_{\mathbf{k}} |0\rangle = 0, and the singly occupied state is simply |1_{\mathbf{k}}\rangle = a_{\mathbf{k}}^\dagger |0\rangle. This structure guarantees antisymmetric wavefunctions and was introduced to handle systems obeying . For multi-mode systems, the full Fock space is built as a tensor product over all single-particle modes, with basis states labeled by the set of occupation numbers \{n_{\mathbf{k}}\} across the modes. The total particle number operator is \hat{N} = \sum_{\mathbf{k}} \hat{n}_{\mathbf{k}}, where \hat{n}_{\mathbf{k}} = a_{\mathbf{k}}^\dagger a_{\mathbf{k}} counts particles in mode \mathbf{k}, and \hat{N} commutes with the Hamiltonian for number-conserving interactions. This occupation number representation simplifies calculations in many-body physics; for instance, the ground state of free fermions, termed the , fills all modes up to the with n_{\mathbf{k}} = 1 for |\mathbf{k}| \leq k_F and n_{\mathbf{k}} = 0 otherwise, yielding a total particle number N = \sum_{\mathbf{k} \leq k_F} 1.

Parastatistics and generalizations

Parastatistics represents a generalization of the standard Bose-Einstein and Fermi-Dirac statistics for indistinguishable particles, allowing for symmetries beyond the simple symmetric or antisymmetric wavefunctions. Proposed by H. S. Green in 1953, this framework describes particles that transform according to higher-dimensional irreducible representations of the permutation group S_N for N particles, rather than the one-dimensional representations corresponding to bosons and fermions. Specifically, para-bosons of order p (where p is a positive integer) are associated with representations having at most p rows in their Young tableaux, while para-fermions of order q have at most q columns; the standard cases recover bosons for p=1 and fermions for q=1. In the second quantization formalism, parastatistics is realized through trilinear commutation relations for the creation (a^\dagger) and annihilation (a) operators, extending the bilinear relations of ordinary particles. For para-bosons, these include relations such as [[a_k, a_l^\dagger], a_m] = 2 \delta_{lm} a_k, where the double commutator captures the higher-order symmetry, differing from the simple [a_k, a_l^\dagger] = \delta_{kl} for bosons. Similar trilinear anticommutation relations apply to para-fermions. These algebraic structures ensure consistency with quantum field theory principles while permitting occupation numbers up to the order parameter, such as at most p particles per state for para-bosons. Parastatistics offers a broader class of statistics, alongside other generalizations such as the fractional statistics of anyons in two dimensions, but extends to higher symmetries in three or more dimensions. Although no experimental evidence for paraparticles has been observed, the theory has been explored for exotic matter, such as in models of dense nuclear matter or beyond-standard-model physics. As of 2025, while no experimental evidence has been observed, theoretical proposals have demonstrated that non-trivial parastatistics can emerge as quasiparticle excitations in solvable quantum spin models in one and two dimensions. Notably, in the context of quark models, O. W. Greenberg proposed para-fermi statistics of order 3 in 1964 to reconcile the Pauli exclusion principle with baryons composed of three quarks sharing quantum numbers, linking it to an SU(3) color-like symmetry; however, the Standard Model ultimately employs ordinary Fermi statistics for quarks with SU(3) color as a gauge degree of freedom rather than parastatistics.

Observables and Measurements

Measuring indistinguishable particles

In quantum mechanics, measurements on indistinguishable particles require observables that are invariant under particle permutations to respect the underlying symmetry of the Hilbert space. For identical particles, whether bosons or fermions, the total wavefunction resides in a symmetrized (or antisymmetrized) subspace, so physical observables must commute with the exchange operator to yield consistent, permutation-invariant results./08%3A_Multiparticle_Systems/8.01%3A_Distinguishable_and_Indistinguishable_Particles) Individual particle positions or momenta cannot be directly measured in a way that distinguishes one particle from another, as such operators would break the permutation symmetry; instead, measurable quantities include one-particle density operators or the total momentum of the system. The one-particle density, for instance, provides the probability distribution for finding a particle at a given position, averaged over all identical particles without labeling them. Similarly, the total momentum operator sums contributions from all particles symmetrically, allowing detection of collective motion without identifying specific individuals. Upon measurement, the wavefunction collapses onto an eigenstate within the symmetric subspace, preserving the overall permutation symmetry of the system. This collapse resolves the inherent ambiguity of "which particle" was measured through the statistical properties dictated by the particles' bosonic or fermionic nature, ensuring that outcomes align with the symmetrized description rather than classical labeling. A practical example is the photoelectric effect, where identical electrons are ejected from a metal surface by incident photons; the measured photocurrent reflects the collective ejection rate and energy distribution of these indistinguishable fermions, rather than tracking individual electron trajectories. This collective observable captures the statistical behavior enforced by the antisymmetric wavefunction, highlighting how indistinguishability manifests in macroscopic currents./08%3A_Multiparticle_Systems/8.01%3A_Distinguishable_and_Indistinguishable_Particles) The no-cloning theorem further underscores limitations in measuring and manipulating indistinguishable particles, prohibiting the creation of identical copies of an unknown quantum state within the symmetrized space, which prevents perfect replication of individual particle states without disturbing the collective symmetry. This implication holds fundamentally for identical particles, where attempts at cloning would violate the linearity of quantum evolution in the permutation-invariant framework.

Identical particle effects in scattering

In scattering processes involving identical particles, the indistinguishability requires the total wave function to be symmetrized or antisymmetrized, leading to interference between direct and exchange amplitudes. For two identical particles, the scattering amplitude includes contributions from the direct process, where particle 1 scatters to angle θ and particle 2 to π - θ, and the exchange process, where the roles are swapped. This results in a total amplitude of the form f(θ) + ε f(π - θ), where ε = +1 for bosons (symmetric wave function) and ε = -1 for fermions (antisymmetric wave function), assuming no phase difference for simplicity. The differential cross section then arises from the modulus squared of this amplitude, given by |f_dir + ε f_exch|^2, where f_dir and f_exch represent the direct and exchange scattering amplitudes, respectively. For bosons, the constructive interference (ε = +1) enhances the cross section, particularly at forward and backward angles where θ ≈ π - θ, while for fermions, the destructive interference (ε = -1) suppresses it, and the cross section vanishes at θ = 90° for spinless cases. This exchange contribution is crucial in distinguishing identical particle effects from classical distinguishable scattering. In electron-electron scattering within metals, these effects manifest in the inelastic scattering rates, where exchange corrections modify the cross sections by 10-20% for incident energies around 1-10 eV, influencing electron transport and resistivity. The antisymmetric nature for these fermions leads to reduced scattering probabilities in certain spin channels, such as the singlet state, compared to the triplet. Similarly, in particle physics, pion-pion scattering treats pions as identical bosons, requiring symmetrization of the amplitude; this exchange symmetry is incorporated in models like , enhancing low-energy cross sections due to constructive interference and aiding in the determination of pion decay constants. Partial wave analysis for identical particles must account for the symmetry by restricting the allowed angular momenta l. For spinless bosons, only even partial waves (even l) contribute to maintain overall symmetry under particle exchange, while for spinless fermions, only odd partial waves (odd l) are permitted. This adjustment ensures the wave function satisfies the required statistics, altering phase shifts and scattering lengths compared to distinguishable cases; for example, s-wave (l=0) scattering is forbidden for fermions but allowed for bosons.

Statistical Consequences

Indistinguishability in statistical mechanics

In classical statistical mechanics, the treatment of indistinguishable particles requires a correction to the phase space volume to account for the overcounting of permutations among identical particles. For a system of N indistinguishable particles, the configurational integral or the total number of microstates is divided by N! to avoid distinguishing particles that are inherently identical, ensuring that thermodynamic quantities like entropy remain extensive. This adjustment resolves the Gibbs paradox, where mixing two identical ideal gases would otherwise predict a spurious entropy increase due to the naive counting of distinguishable states; with the $1/N! factor, no such entropy of mixing occurs for identical gases, aligning with thermodynamic expectations. In the quantum mechanical approach, indistinguishability is incorporated by restricting the Hilbert space to symmetric subspaces for bosons or antisymmetric subspaces for fermions, reflecting the exchange symmetry of the wave function. The partition function Z for such a system is then computed as the trace over this appropriate subspace: Z = \operatorname{Tr} \left( e^{-\beta H} \right), where \beta = 1/(k_B T), H is the Hamiltonian, and the trace enforces the permutation invariance of the states. This quantum route naturally emerges from the symmetrization postulate and avoids the ad hoc $1/N! correction of the classical limit by projecting onto the physically allowed representations of the permutation group. For an ideal monatomic gas, the entropy expression derived from this framework is the Sackur-Tetrode equation, which explicitly includes the $1/N! factor to correct for indistinguishability while incorporating quantum phase space considerations via Planck's constant h. The equation takes the form S = N k_B \left[ \ln \left( \frac{V}{N} \left( \frac{4\pi m U}{3 N h^2} \right)^{3/2} \right) + \frac{5}{2} \right], where V is the volume, m the particle mass, U the internal energy, and k_B Boltzmann's constant; this yields an entropy that is extensive and resolves inconsistencies in the classical ideal gas entropy for large N. The derivation starts from the microcanonical multiplicity \Omega \approx V^N (2\pi m U)^{3N/2} / (h^{3N} N! (3N/2)!), applying Stirling's approximation to obtain the final form, with $1/N! preventing overcounting of identical particle configurations. Indistinguishability also affects fluctuations in observables, such as particle number in a given state. For bosons, quantum statistics lead to enhanced number fluctuations compared to the classical Maxwell-Boltzmann case, with the relative variance \left( \langle \hat{n}_r^2 \rangle - \langle \hat{n}_r \rangle^2 \right) / \langle \hat{n}_r \rangle^2 = 1/\langle \hat{n}_r \rangle + 1, reflecting bunching tendencies. In contrast, fermions exhibit suppressed fluctuations due to the , with relative variance $1/\langle \hat{n}_r \rangle - 1, smaller than the classical value of $1/\langle \hat{n}_r \rangle. These differences arise from the occupation number statistics in the and highlight how exchange symmetry modifies statistical weights beyond classical predictions.

Bose-Einstein and Fermi-Dirac distributions

In the grand canonical ensemble, which allows for fluctuations in particle number while fixing the average via the chemical potential μ, the statistical distributions for indistinguishable particles arise from the symmetry of their wave functions under exchange. For bosons, the symmetric wave function permits multiple occupations of the same state, while for fermions, the antisymmetric wave function enforces the Pauli exclusion principle, limiting each state to at most one particle. The average occupation number ⟨n_k⟩ for a single-particle state with energy ε_k is derived by evaluating the grand partition function Ξ = ∑_{n=0}^∞ e^{-β n (ε_k - μ)} for bosons (summing over n = 0,1,2,...), yielding ⟨n_k⟩ = 1 / (e^{β(ε_k - μ)} - 1), where β = 1/(k_B T) and k_B is Boltzmann's constant. For fermions, the sum is restricted to n = 0 or 1, giving ⟨n_k⟩ = 1 / (e^{β(ε_k - μ)} + 1). These expressions, known as the Bose-Einstein and Fermi-Dirac distributions, respectively, were first derived in the context of quantum statistics. An alternative derivation maximizes the entropy subject to constraints on average energy and particle number, incorporating the appropriate symmetry: for bosons, the multiplicity of states is ∏_k (n_k + g_k - 1 choose n_k) where g_k is the degeneracy, leading to the upon using ; for fermions, it is ∏_k (g_k choose n_k), yielding the . Using second quantization, the occupation number operator â†_k â_k for bosons (or ĉ†_k ĉ_k for fermions) has eigenvalues following these distributions in thermal equilibrium, as the grand potential Φ = -k_B T ln Ξ directly gives the averages. The chemical potential μ is determined self-consistently from the total average particle number N = ∑_k ⟨n_k⟩, ensuring the distributions describe systems with fixed average density. For the Bose-Einstein distribution, μ must satisfy μ < min{ε_k} to ensure positive occupation numbers and convergence of the partition function, allowing ⟨n_k⟩ > 1 for low-energy states. At low , when the thermal de Broglie wavelength exceeds the interparticle spacing, a macroscopic fraction of particles can occupy the (ε_0 ≈ 0), with μ approaching 0 from below; this phenomenon, Bose-Einstein condensation, occurs below a critical temperature T_c ∝ (N/V)^{2/3} h^2 / (m k_B), marking a to a coherent . Einstein predicted this condensation for ideal gases in his extension of Bose's photon statistics to massive particles. In contrast, the Fermi-Dirac distribution enforces ⟨n_k⟩ ≤ 1 due to the +1 in the denominator, reflecting fermionic exclusion. At (T=0), μ equals the E_F, filling all states up to E_F and leaving higher states empty, forming a sharp ; the system behaves as a degenerate with zero-point pressure supporting white dwarfs. For T > 0, thermal excitations smear the distribution over ~k_B T around E_F, but the exclusion principle dominates at low T, leading to distinct thermodynamic properties like linear specific heat. Fermi introduced this statistics for electrons in metals, independently developed by Dirac in the context of .

Topological Aspects

Homotopy classes in particle exchange

The configuration space for a system of N indistinguishable particles in three-dimensional Euclidean space is defined as the quotient space C_N = \left( (\mathbb{R}^3)^N \setminus \Delta \right) / S_N, where \Delta is the set of diagonal points where any two particles coincide, and S_N is the symmetric group acting by permuting particle labels. This space captures the indistinguishability by identifying configurations that differ only by particle relabeling, excluding coincidences to avoid singularities. The classes of paths in this configuration space, classified by its \pi_1(C_N), determine the possible exchange statistics for the particles. In three dimensions, since \mathbb{R}^3 is simply connected, \pi_1(C_N) \cong S_N. The elements of S_N correspond to permutations induced by continuous exchanges of particle positions, and the one-dimensional unitary representations of S_N yield only two possibilities: the trivial representation, giving +1 upon (bosons), or the alternating representation, giving -1 (fermions). Particle exchanges are represented by closed paths in C_N, known as braids in the covering space, but in three dimensions, these paths are homotopy invariant only up to the discrete structure of S_N. A single exchange of two particles generates a transposition in S_N, while a double exchange is homotopic to the identity, reinforcing the \pm 1 phase structure without intermediate possibilities. This topological framework connects to the spin-statistics theorem, which dictates that particles with integer spin obey bosonic statistics (even permutations, phase +1), while half-integer spin particles obey fermionic statistics (odd permutations, phase -1). The theorem arises because the total wave function, incorporating orbital and spin parts, must transform consistently under exchanges, with the spatial homotopy classes tying directly to the spin representation's behavior under 360-degree rotations.

Anyons and fractional statistics

In two spatial dimensions, the configuration space of indistinguishable particles allows for more complex exchange paths compared to higher dimensions, where the fundamental group of the configuration space is the B_N for N particles. This structure permits abelian representations of exchanges, resulting in a e^{i \theta} upon interchanging two particles, where \theta can take any real value rather than being restricted to integer multiples of \pi. Particles exhibiting such fractional exchange phases are known as anyons, a term coined by in 1982 to describe quasiparticles whose statistics interpolate continuously between bosonic and fermionic behaviors. For bosons, \theta = 0, yielding no change upon ; for fermions, \theta = \pi, producing the familiar antisymmetric . Intermediate cases include semions with \theta = \pi/2, which acquire a of i under and have been proposed in models of . More exotic non-abelian anyons, where braiding leads to unitary transformations in a degenerate rather than simple phases, are of particular interest for topological due to their inherent against decoherence. A prominent physical realization of anyons occurs in the (FQHE), where Laughlin in two-dimensional electron gases under strong magnetic fields at filling factor \nu = 1/m (with m odd) exhibit fractional charge e/m and statistics parameterized by \theta = \pi / m. These emerge as excitations in the Laughlin wave function, and their fractional statistics arises from the topological properties of the ground state, as deduced via the adiabatic Berry accumulated during exchanges. Early experiments in the , such as resonant tunneling measurements, confirmed the fractional charge and implied the associated statistics. Direct observation of anyonic braiding statistics was achieved in 2020 using in the \nu = 1/3 , providing definitive evidence. In general, anyonic statistics can be characterized by a \alpha = \theta / \pi, which smoothly interpolates between (\alpha = 0) and Fermi (\alpha = 1) limits, enabling a rich variety of quantum phases in two-dimensional systems. In three dimensions, the homotopy classes of particle exchanges reduce to the , limiting \theta to 0 or \pi (modulo $2\pi) and recovering standard Bose-Fermi statistics as a special case.

References

  1. [1]
    [PDF] Chapter 8 - Identical Particles - MIT OpenCourseWare
    May 8, 2022 · Of course, two identical particles can have different momentum, energy, angular momentum. For example all electrons are identical, all protons, ...
  2. [2]
  3. [3]
    [PDF] Discerned and Non-Discerned Particles in Classical Mechanics and ...
    Jul 29, 2010 · Definition 1 - A classical particle is potentially discerned if its initial position xo and its initial velocity v0 are known. Let us note that ...
  4. [4]
    Distribution functions for identical particles - HyperPhysics
    The Maxwell-Boltzmann distribution is the classical distribution function for distribution of an amount of energy between identical but distinguishable ...
  5. [5]
    [PDF] arXiv:0903.4748v2 [physics.chem-ph] 29 Jun 2009
    Jun 29, 2009 · The inextensivity leads to the so-called Gibbs paradox in which the mixing entropy of two identical classical gases increases.Missing: early | Show results with:early
  6. [6]
    The Gibbs Paradox: Lessons from Thermodynamics - PMC - NIH
    The Gibbs paradox in statistical mechanics is often taken to indicate that already in the classical domain particles should be treated as fundamentally ...
  7. [7]
    The Uncertainty Principle (Stanford Encyclopedia of Philosophy)
    Oct 8, 2001 · The uncertainty principle (for position and momentum) states that one cannot assign exact simultaneous values to the position and momentum of a physical system.
  8. [8]
    wave-particle duality, uncertainity principle - University of Oregon
    The uncertainty principle, developed by W. Heisenberg, is a statement of the effects of wave-particle duality on the properties of subatomic objects. Consider ...
  9. [9]
    [PDF] Bose-Einstein Statistics - The Information Philosopher
    Einstein's most profound insight into elementary particles might be their indistinguishability, their interchangeability. Particles are not independent of one ...
  10. [10]
    [PDF] The concept 'indistinguishable'1 Simon Saunders - arXiv
    Fairly directly, and uncontrovertibly, that takes us to a series of papers by Einstein on the quantum theory of the ideal monatomic gas, published in 1924-25 ( ...
  11. [11]
  12. [12]
  13. [13]
    Helium Energy Levels - HyperPhysics
    The helium ground state consists of two identical 1s electrons. The energy required to remove one of them is the highest ionization energy of any atom in the ...
  14. [14]
  15. [15]
    [PDF] Indistinguishable Particles in Quantum Mechanics - arXiv
    Nov 1, 2005 · The Bose-Einstein distribution was first proposed by Satyendranath Bose in 1924 for a gas of photons [13] and very soon after generalized by ...
  16. [16]
    [PDF] Phys 487 Discussion 1 – Identical Particles
    If you place two electrons in a box, you have a system with two identical particles. ... 2-Boson wavefunctions are Symmetric under exchange → ψ ( ! r2, ! r1) = +ψ ...
  17. [17]
    [PDF] Chapter 4: Identical Particles - MIT OpenCourseWare
    To describe indistinguishable particles, we can use set notation. ... "Supplementary Notes: Canonical Quantization and. Application to the Quantum Mechanics of a ...
  18. [18]
    [PDF] 1 He atom and other atoms with more than two electrons - bingweb
    J is the exchange integral. (units of energy) and is related to the overlap of the charge distributions of atoms i and j. i. ˆ. S is the spin operator ...<|separator|>
  19. [19]
    Helium Atom - Richard Fitzpatrick
    Incidentally, helium in the spin singlet state is known as para-helium, whereas helium in the triplet state is called ortho-helium. As we have seen, for the ...
  20. [20]
    On the theory of quantum mechanics - Journals
    Abstract. 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 ...
  21. [21]
    [2106.14679] Permanent variational wave functions for bosons - arXiv
    Jun 25, 2021 · Permanent states, the bosonic counterpart of Slater determinants, are constructed by symmetrizing a product of single-particle orbitals and are ...
  22. [22]
    [PDF] An Introduction to Hartree-Fock Molecular Orbital Theory
    Again, the Hartree-Fock method seeks to approximately solve the electronic Schrödinger equation, and it assumes that the wavefunction can be approximated by a ...
  23. [23]
  24. [24]
    Physicists describe exotic 'paraparticles' that defy categorization
    Jan 8, 2025 · ... electrons but to a broader class of particles, including protons and neutrons, which they called fermions. Conversely, particles that do ...
  25. [25]
    The quantum theory of the emission and absorption of radiation
    The new quantum theory, based on the assumption that the dynamical variables do not obey the commutative law of multiplication, has by now been developed ...
  26. [26]
    Konfigurationsraum und zweite Quantelung | Zeitschrift für Physik A ...
    Fock, V. Konfigurationsraum und zweite Quantelung. Z. Physik 75, 622–647 (1932). https://doi.org/10.1007/BF01344458. Download citation. Received: 10 March 1932.
  27. [27]
    Über das Paulische Äquivalenzverbot | Zeitschrift für Physik A ...
    Download PDF · Zeitschrift für Physik. Über ... Cite this article. Jordan, P., Wigner, E. Über das Paulische Äquivalenzverbot. Z. Physik 47, 631–651 (1928).
  28. [28]
    [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 ...
  29. [29]
    Second quantization in nonrelativistic quantum mechanics
    The commutation (anticommutation) rela- tions for boson (fermion) field creation and annihilation operators are derived from the symmetry (antisymmetry) of the ...Missing: seminal | Show results with:seminal
  30. [30]
    [PDF] Chapter 4: Identical Particles
    Jan 4, 2024 · In a quantum field theory, the position serves as an index for the various field operators, and is not itself an observable. Eqs. (4.125) and ( ...
  31. [31]
    Identical Quantum Particles, Entanglement, and Individuality - PMC
    Particles in classical physics are distinguishable objects, which can be picked out individually on the basis of their unique physical properties.
  32. [32]
    [PDF] arXiv:2011.08839v2 [quant-ph] 29 May 2021
    May 29, 2021 · We show that symmetric measurements performed on identical particles can in principle discriminate between the product and symmetrized states.
  33. [33]
    A single quantum cannot be cloned - Nature
    Oct 28, 1982 · We show here that the linearity of quantum mechanics forbids such replication and that this conclusion holds for all quantum systems.
  34. [34]
    [PDF] arXiv:2102.00780v1 [quant-ph] 1 Feb 2021
    Feb 1, 2021 · The significance of our result is that for indistinguishable systems, the no-cloning theorem remains more fundamental than. MoE and the former ...
  35. [35]
    The Feynman Lectures on Physics Vol. III Ch. 4: Identical Particles
    If a process involves two particles that are identical, reversing which one arrives at a counter is an alternative which cannot be distinguished.
  36. [36]
    [PDF] Scattering of identical particles - bingweb
    If the particles which are scattered are identical particles, we have to take into account the symmetry properties of the total wave functions, symmetric ...
  37. [37]
    Exchange corrections for inelastic electron scattering rates in ...
    Nov 26, 2019 · The results suggest that the impact of exchange may alter inelastic scattering cross sections by between 10% and 20% for incident energies ...
  38. [38]
    [PDF] Simple Description of Pion-Pion Scattering to 1 GeV
    Nov 16, 1995 · In the large Nc picture the leading amplitude (of order 1/Nc) is a sum of polynomial contact terms and tree type resonance exchanges.
  39. [39]
    [PDF] Scattering Theory III 1 Partial Wave Analysis - 221B Lecture Notes
    If they have spins, their spins need to be interchanged at the same time. If two particles are idential spinless bosons, say two Helium atoms (as- suming. 4.
  40. [40]
    [PDF] Lectures on scattering theory in partial-wave amplitudes - arXiv
    Sep 25, 2024 · [7,9]. ○ Identical particles, unitary normalization, and including isospin: For the case of two bosons/fermions the state must be symmetric/ ...
  41. [41]
    [PDF] THE GIBBS PARADOX - DAMTP
    In the present note we consider the \Gibbs Paradox" about entropy of mixing and the logically inseparable topics of reversibility and the extensive property of ...
  42. [42]
    [PDF] arXiv:2206.12639v2 [quant-ph] 12 Jul 2023
    Jul 12, 2023 · Quantum systems of indistinguishable particles are commonly described using the formalism of second quantisation, which relies on the assumption ...
  43. [43]
    None
    ### Summary of Sackur-Tetrode Equation Derivation (Incorporating 1/N! Factor)
  44. [44]
    [PDF] Chapter 12 Quantum gases
    The particle fluctuations for bosons are larger than those for classical particles while for fermions they are smaller. This result reflect the intuition that ...<|control11|><|separator|>
  45. [45]
    [PDF] Planck's Law and Light Quantum Hypothesis.
    Planck's Law and Light Quantum Hypothesis. S.N. Bose. (Received 1924). Planck's formula for the distribution of energy in the radiation from a black body ...
  46. [46]
    [PDF] Quantum Theory of a Monoatomic Ideal Gas A translation of ...
    Jan 9, 2015 · In this paper Einstein develops for the first time the Bose-Einstein statistics, applying the ideas first discussed in a paper by D. Bose and ...
  47. [47]
    [PDF] On Quantizing an Ideal Monatomic Gas - Gilles Montambaux
    We are therefore forced to assume that the motion of a molecule in an ideal gas is quantized; this quantization manifests itself for low temperatures by certain ...
  48. [48]
    [PDF] On a probabilistic derivation of the basic particle statistics (Bose ...
    Apr 7, 2020 · For k = 1, (3) follows directly. which is the Bose-Einstein distribution. In the next section we derive (8)-(9) from (3) and basic ...<|separator|>
  49. [49]