Fact-checked by Grok 2 weeks ago

Free particle

In physics, a free particle is a model for a particle subject to no external forces or potentials, enabling it to propagate indefinitely without acceleration or confinement. In , a free particle obeys Newton's , maintaining constant and in a field-free , with its given by E = \frac{p^2}{2m}, where p is and m is . This idealized scenario serves as a foundational reference for understanding more complex dynamics under forces. In quantum mechanics, the free particle is described by the time-independent Schrödinger equation -\frac{\hbar^2}{2m} \frac{d^2 \psi(x)}{dx^2} = E \psi(x), where \psi(x) is the wave function, \hbar is the reduced Planck's constant, and E is energy. The solutions are plane waves of the form \psi_{\pm}(x) = A_{\pm} e^{\pm i k x}, with wave number k = \sqrt{2mE}/\hbar, corresponding to definite momenta p = \pm \hbar k and continuous energy spectra due to the absence of boundaries. These states exhibit uniform probability density over all space, illustrating the Heisenberg uncertainty principle: precise momentum knowledge implies complete positional delocalization. In relativistic quantum mechanics, free particles are described by wave equations such as the Klein–Gordon equation for spin-0 particles and the Dirac equation for spin-1/2 particles. The free particle model is crucial for deriving approximations in scattering theory, solid-state physics, and understanding wave propagation in unbounded systems.

Classical Mechanics

Definition and Kinematics

In classical mechanics, a free particle is defined as a point mass that experiences no net external force acting upon it. According to Newton's of motion, such a particle will continue in a state of rest or uniform motion in a straight line at constant unless compelled to change its state by forces impressed upon it. This idealization assumes an absence of interactions with other bodies, fields, or constraints, allowing the particle's trajectory to be purely inertial. The of a free particle are straightforward and deterministic, governed solely by its initial conditions. The of the particle as a of time is given by \mathbf{r}(t) = \mathbf{r}_0 + \mathbf{v} t, where \mathbf{r}_0 is the initial vector at t = 0, and \mathbf{v} is the constant vector, of time. This linear implies zero (\mathbf{a} = d\mathbf{v}/dt = 0), and the speed |\mathbf{v}| remains unchanged. In one dimension, for simplicity, this reduces to x(t) = x_0 + v t, highlighting the uniform progression along the path. This behavior is fundamentally tied to the concept of inertial reference frames, where the laws of motion hold without fictitious forces. In such frames, the free particle's motion exhibits : its velocity appears constant regardless of the frame's uniform translation relative to another inertial frame. This invariance underpins the relativity of uniform motion in non-relativistic physics, distinguishing free particles from those under acceleration. Examples of free particle motion include the idealized of a far from gravitational influences in interplanetary space or an air in a perfect vacuum, where collisions are negligible. In contrast, real particles are often constrained by potentials, such as gravitational or electromagnetic fields, leading to curved paths—unlike the unbound straight-line propagation of the truly free case.

Dynamics and Conservation Laws

The dynamics of a classical free particle are governed by the absence of external forces, leading directly to Newton's of motion, which states that a particle remains at rest or in uniform rectilinear motion unless acted upon by a net external . This law establishes that the of the particle is zero, \mathbf{a} = 0, implying constant \mathbf{v}. Due to in space—the invariance of the system's laws under spatial displacements—the linear \mathbf{p} = m \mathbf{v} of the free particle is conserved, remaining constant throughout its motion. This conservation arises from , which links continuous of the action to conserved quantities; for a free particle, the under translations \delta \mathbf{x} = \epsilon yields the \frac{d\mathbf{p}}{dt} = 0. Similarly, time-translation invariance implies the conservation of total energy. For a free particle, the E = \frac{1}{2} m v^2 is the sole contribution and remains constant, as there are no non-conservative forces to alter it. This follows from the time-independence of the , ensuring \frac{dE}{dt} = 0 via the Euler-Lagrange . In the formulation, the free particle's is L = \frac{1}{2} m \dot{x}^2 (in one dimension for ), where the serves as the since vanishes. Applying the Euler-Lagrange equation \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{x}} \right) - \frac{\partial L}{\partial x} = 0 gives \frac{d}{dt} (m \dot{x}) = 0, confirming constant \dot{x} = v. The Hamiltonian formulation provides an equivalent description, with the Hamiltonian H = \frac{p^2}{2m} expressed in terms of momentum p = m \dot{x}. Hamilton's equations, \dot{x} = \frac{\partial H}{\partial p} = \frac{p}{m} and \dot{p} = -\frac{\partial H}{\partial x} = 0, reproduce the constant velocity and conserved momentum, while the time-independence of H underscores energy conservation.

Non-Relativistic Quantum Mechanics

Schrödinger Equation Formulation

In 1926, introduced the wave equation as part of his formulation of wave mechanics, with the free particle representing the simplest case where no external potential acts on the system. The derivation of the for a free particle proceeds from the classical via rules. The classical Hamiltonian for a free particle of mass m is H = \frac{p^2}{2m}, where p is the . In , position x and p are promoted to operators satisfying the commutation relation [x, p] = i\hbar, with p = -i\hbar \nabla in the position representation. Substituting into the quantum yields the operator \hat{H} = -\frac{\hbar^2}{2m} \nabla^2. The time evolution of the wave function \psi(\mathbf{r}, t) follows from the correspondence principle, where the classical is replaced by the , leading to the time-dependent : i\hbar \frac{\partial \psi}{\partial t} = -\frac{\hbar^2}{2m} \nabla^2 \psi. For stationary states, where the wave function separates as \psi(\mathbf{r}, t) = \phi(\mathbf{r}) e^{-iEt/\hbar}, the time-dependent equation reduces to the time-independent form: -\frac{\hbar^2}{2m} \nabla^2 \phi = E \phi, with E as the energy eigenvalue. This eigenvalue equation describes the spatial behavior of the particle in energy eigenstates. The free particle exists in infinite space with zero potential V = 0 everywhere, imposing no confining boundary conditions. Consequently, the domain of the wave function extends over all space, \mathbf{r} \in \mathbb{R}^3, without restrictions on the wave function at finite boundaries. The probabilistic interpretation, proposed by in 1926, assigns |\psi(\mathbf{r}, t)|^2 as the probability density for finding the particle at position \mathbf{r} at time t. For the equation to yield physically meaningful probabilities, the wave function must satisfy the normalization condition \int |\psi|^2 d^3\mathbf{r} = 1. However, in infinite space, exact energy eigenstates are plane waves that extend indefinitely, rendering them non-normalizable as their integral diverges; practical treatments often employ wave packets or impose in a large but finite volume to approximate normalization.

Plane Wave Solutions

In non-relativistic , the plane wave solutions to the free particle take the form \psi(x,t) = A \exp\left[i (k x - \omega t)\right], where A is a constant , k is the wave number, and \omega is the . These functions satisfy the time-dependent i \hbar \frac{\partial \psi}{\partial t} = -\frac{\hbar^2}{2m} \frac{\partial^2 \psi}{\partial x^2} for zero potential. Substituting the plane wave into the yields the \omega = \frac{\hbar k^2}{2m}, linking the to the wave number and particle mass m. This relation follows from the eigensolution property of the equation. The plane waves serve as eigenfunctions of the momentum operator \hat{p} = -i \hbar \frac{\partial}{\partial x}, with eigenvalue p = \hbar k. The associated energy eigenvalue is E = \hbar \omega = \frac{p^2}{2m}, reflecting the classical kinetic energy in quantum form. Plane waves extend infinitely in space and thus are not square-integrable, preventing normalization in the usual L^2 sense over infinite domains. Instead, they employ Dirac delta function normalization, where the position representation satisfies \int_{-\infty}^{\infty} \psi_p^*(x) \psi_{p'}(x) \, dx = \delta(p - p'), ensuring orthogonality for different momenta. By the superposition principle, any general solution to the free particle Schrödinger equation can be expressed as an integral over plane waves: \psi(x,t) = \frac{1}{\sqrt{2\pi}} \int_{-\infty}^{\infty} \tilde{\psi}(k) \exp\left[i (k x - \omega(k) t)\right] \, dk, with \tilde{\psi}(k) as the Fourier amplitude and \omega(k) = \frac{\hbar k^2}{2m}. This Fourier decomposition expands arbitrary initial wave functions in the continuous momentum basis. Physically, plane waves describe delocalized states with precise momentum p = \hbar k but infinite uncertainty in position, embodying the trade-off in the Heisenberg uncertainty principle \Delta x \Delta p \geq \hbar/2. Such states are idealized, representing particles with uniform probability density across all space.

Wave Packets and Superpositions

In , plane waves provide exact solutions to the free particle but are delocalized over all space, failing to represent a particle confined to a finite region. To model a localized particle with an approximate and , the wave function is constructed as a superposition of plane waves with a continuous range of wave numbers centered around a mean value. This approach, introduced by Schrödinger as "wave groups," allows the probability density to be concentrated in a small spatial region while satisfying the of the . The general form of such a in one dimension is given by the representation \psi(x,t) = \int_{-\infty}^{\infty} \mathrm{d}k \, \phi(k) \, \exp\left[i \left( k x - \omega(k) t \right)\right], where \phi(k) is the in momentum space, serving as the of the initial , and \omega(k) = \hbar k^2 / (2m) is the for a free particle of m. The \phi(k) is typically chosen to be peaked around some central wave number k_0, ensuring a spread in \Delta k that corresponds to a spatial localization via the \Delta x \Delta p \gtrsim \hbar/2. This representation normalizes the such that \int |\psi(x,t)|^2 \mathrm{d}x = 1, provided \int |\phi(k)|^2 \mathrm{d}k = 1. A common example is the Gaussian wave packet, which minimizes the position-momentum uncertainty and provides an analytically tractable form. At t=0, it takes the form \psi(x,0) = \left( \frac{1}{2\pi \sigma^2} \right)^{1/4} \exp\left[ -\frac{x^2}{4\sigma^2} + i k_0 x \right], where \sigma is the initial spatial width. The corresponding \phi(k) is also Gaussian, centered at k_0 with width $1/(2\sigma). This initial state is localized around the mean position \langle x \rangle = 0 (or shifted accordingly), with mean momentum \langle p \rangle = \hbar k_0, as computed from the expectation values \langle x \rangle = \int \psi^* x \psi \, \mathrm{d}x and \langle p \rangle = -i\hbar \int \psi^* \frac{\partial \psi}{\partial x} \, \mathrm{d}x. Such wave packets bridge classical and quantum descriptions by having their center of probability density follow a classical under the , \frac{\mathrm{d} \langle x \rangle}{\mathrm{d}t} = \langle p \rangle / m and \frac{\mathrm{d} \langle p \rangle}{\mathrm{d}t} = -\left\langle \frac{\partial V}{\partial x} \right\rangle, which for a free particle (V=0) yields uniform motion. However, since the underlying plane waves extend infinitely, constructing normalizable wave packets in unbounded space demands careful choice of \phi(k) with finite support or rapid decay to ensure of the and proper .

Group and Phase Velocities

In non-relativistic quantum mechanics, the free particle is described by plane wave solutions to the Schrödinger equation, which exhibit a dispersion relation \omega(k) = \frac{\hbar k^2}{2m}, where \omega is the angular frequency, k is the wave number, \hbar is the reduced Planck's constant, and m is the particle mass. This relation arises directly from substituting the plane wave form \psi(x,t) = e^{i(kx - \omega t)} into the time-dependent Schrödinger equation for zero potential, i\hbar \frac{\partial \psi}{\partial t} = -\frac{\hbar^2}{2m} \frac{\partial^2 \psi}{\partial x^2}. The v_p of these waves is defined as the speed at which a constant phase point propagates, given by v_p = \frac{\omega}{k} = \frac{\hbar k}{2m}. This velocity depends linearly on k, meaning it varies for different components and does not correspond to the motion of a classical particle. In contrast, the v_g, which represents the propagation speed of the overall wave envelope, is obtained by differentiating the : v_g = \frac{d\omega}{dk} = \frac{\hbar k}{m}. Since the p = \hbar k, this simplifies to v_g = \frac{p}{m}, matching the classical of a free particle with the same . Physically, the phase velocity characterizes the motion of individual components, while the describes the velocity of the wave packet's envelope, which localizes the particle's probable position and thus aligns with the classical particle trajectory. For a wave packet constructed as a superposition of plane waves with a narrow distribution of k values around a central k_0, the approximates the classical v = \frac{\hbar k_0}{m}, providing a clear quantum analog to particle motion.

Wave Packet Spreading and Uncertainty

In , the wave function of a free particle, initially localized as a , undergoes temporal evolution governed by the , leading to a gradual broadening of the packet even in the absence of any potential. This spreading arises from the in the components of the superposition, where plane waves with different wavelengths propagate at different velocities. For a Gaussian wave packet, which achieves the minimum in and , the time-dependent width \sigma(t) is given by \sigma(t)^2 = \sigma(0)^2 + \left( \frac{\hbar t}{2 m \sigma(0)} \right)^2, where \sigma(0) is the initial width, m is the particle mass, \hbar is the reduced Planck's constant, and t is time. The derivation of this formula proceeds from the time evolution of the position variance, defined as \langle x^2 \rangle - \langle x \rangle^2. For a free particle, the expectation value \langle x \rangle moves with constant group velocity, but the second moment \langle x^2 \rangle increases due to the spread in velocities from the momentum distribution. Using the Fourier transform of the initial Gaussian wave function and applying the time-dependent phase factors e^{-i E(k) t / \hbar} with E(k) = \hbar^2 k^2 / (2m), the evolved wave function remains Gaussian, allowing exact computation of the moments via Gaussian integrals. This yields the quadratic growth in variance, reflecting the dispersive nature of the non-relativistic dispersion relation \omega(k) \propto k^2. This spreading directly illustrates the Heisenberg uncertainty principle, \Delta x \Delta p \geq \hbar / 2, where the initial minimum-uncertainty Gaussian satisfies \Delta x(0) \Delta p = \hbar / 2 with \Delta p = \hbar / (2 \Delta x(0)). As time progresses, the position uncertainty \Delta x(t) \approx \sigma(t) increases while \Delta p remains constant, maintaining the product above the minimum but demonstrating how the fixed momentum spread \Delta p leads to differential velocities \Delta v = \Delta p / m, causing the packet to broaden. A narrower initial localization (smaller \sigma(0)) implies a larger \Delta p, resulting in faster spreading, underscoring the principle's dynamic implications. Physically, this phenomenon implies that quantum particles exhibit diffusive-like behavior, delocalizing over time unlike classical point particles that maintain fixed trajectories without forces. For long times, t \gg 2 m \sigma(0)^2 / \hbar, the width grows linearly as \sigma(t) \approx \hbar t / (2 m \sigma(0)), with the rate inversely proportional to and initial width, making it pronounced for light particles like electrons but negligible for macroscopic objects due to the small scale of \hbar. This spreading highlights the inherently probabilistic nature of quantum propagation for free particles.

Relativistic Quantum Mechanics

Klein-Gordon Equation

The Klein-Gordon equation serves as the fundamental relativistic wave equation for free particles of spin zero, extending the principles of quantum mechanics to incorporate . It arises from the quantization of the classical relativistic energy-momentum relation E^2 = p^2 c^2 + m^2 c^4, where E is the , p the , m the rest , and c the . By promoting E to the i \hbar \partial_t and p to -i \hbar \nabla, and squaring to obtain a second-order equation, the result is the Klein-Gordon equation in its time-dependent form: \left( \frac{1}{c^2} \frac{\partial^2}{\partial t^2} - \nabla^2 + \frac{m^2 c^2}{\hbar^2} \right) \psi(\mathbf{x}, t) = 0, or, in covariant notation using the d'Alembertian operator \square = \partial^\mu \partial_\mu, (\square + \frac{m^2 c^2}{\hbar^2}) \psi(x) = 0. This equation was independently derived by Oskar Klein and Walter Gordon in 1926, predating Paul Dirac's work on spin-1/2 particles, as part of early efforts to reconcile quantum mechanics with relativity. Multiple physicists, including Vladimir Fock, contributed similar derivations that year, reflecting the equation's rapid emergence in the literature. The formulation ensures Lorentz invariance, treating the wave function \psi as a scalar field describing the particle's probability amplitude. Plane wave solutions to the Klein-Gordon equation take the form \psi(\mathbf{x}, t) \propto \exp\left[-i (E t - \mathbf{p} \cdot \mathbf{x})/\hbar \right], where the dispersion relation is E = \pm \sqrt{p^2 c^2 + m^2 c^4}. These solutions include both positive-energy states (corresponding to particles) and negative-energy states (initially interpreted as antiparticles or problematic excitations). Superpositions of such can form wave packets that propagate relativistically, with the v_g = \partial E / \partial p = p c^2 / E matching the classical for positive-energy components. Despite its formal successes, the Klein-Gordon equation encounters significant challenges in single-particle quantum mechanics. The conserved four-current, derived from for the global U(1) phase invariance, yields a \rho = (i \hbar / 2 m c^2) (\psi^* \partial_t \psi - \psi \partial_t \psi^*) that is not positive definite, allowing negative values due to between positive- and negative-energy components. This undermines a probabilistic , as the "probability " can become negative or fail to normalize properly. In the non-relativistic limit (low momenta, p \ll m c), the equation reduces formally to the for positive-energy projections, but the issues persist for low-energy states, where relativistic corrections introduce oscillations and non-positivity not present in the non-relativistic case. These limitations, recognized soon after its proposal, highlighted the need for a second-quantized field-theoretic , where the equation describes a quantum field with particle and creation/annihilation rather than a single-particle .

Dirac Equation Solutions

The Dirac equation provides a relativistic wave equation for free particles with , such as electrons, incorporating both and . It is expressed in Hamiltonian form as i \hbar \frac{\partial \psi}{\partial t} = c \boldsymbol{\alpha} \cdot \mathbf{p} \psi + \beta m c^2 \psi, where \psi is a four-component , \mathbf{p} = -i \hbar \boldsymbol{\nabla} is the , m is the particle , c is the , \hbar is the reduced Planck's constant, and \boldsymbol{\alpha} = (\alpha_x, \alpha_y, \alpha_z) along with \beta are 4×4 Hermitian matrices obeying the anticommutation relations \{\alpha_i, \alpha_j\} = 2\delta_{ij}, \{\alpha_i, \beta\} = 0, and \beta^2 = 1. These matrices ensure the equation is linear and first-order in both space and time derivatives, addressing limitations of prior relativistic formulations like the Klein-Gordon equation, which treated particles as scalars without and led to issues with probability interpretation. Plane wave solutions to the for free particles are of the form \psi(\mathbf{x}, t) = u(\mathbf{p}) \exp\left[-i (E t - \mathbf{p} \cdot \mathbf{x})/\hbar\right] for positive energy states, where u(\mathbf{p}) is a four-component normalized such that \bar{u} u = 2m, and \bar{u} = u^\dagger \beta. For states, the solutions are \psi(\mathbf{x}, t) = v(\mathbf{p}) \exp\left[i (E t - \mathbf{p} \cdot \mathbf{x})/\hbar\right], with v(\mathbf{p}) satisfying \bar{v} v = -2m and interpreted as describing antiparticles, such as positrons, in the picture. Substituting these into the yields (\slash{p} - m) u(\mathbf{p}) = 0 for positive energy and (\slash{p} + m) v(\mathbf{p}) = 0 for , where \slash{p} = \gamma^\mu p_\mu in covariant notation with Dirac matrices \gamma^\mu. The energy spectrum from these solutions is E = \pm \sqrt{p^2 c^2 + m^2 c^4}, where p = |\mathbf{p}|, providing both positive and negative branches that match the relativistic energy-momentum relation. This spectrum resolves the interpretive challenges of negative energy solutions in the Klein-Gordon equation—such as the Klein paradox involving apparent pair production—by reinterpreting them as holes in a filled negative energy sea, corresponding to antiparticles with positive energy. In the massless limit (m \to 0), the Dirac equation reduces to the Weyl equation, with solutions exhibiting two helicity states: particles with spin projection along the momentum direction (+1/2) and opposite (-1/2), reflecting left- and right-handed chiralities that coincide with helicity. Formulated by Paul Dirac in 1928, the equation predicted the existence of antimatter through its negative energy solutions, a prediction experimentally confirmed by the discovery of the positron by Carl D. Anderson in 1932 using cloud chamber observations of cosmic rays.

References

  1. [1]
    5.1: The Free Particle - Chemistry LibreTexts
    Apr 21, 2022 · In classical physics, a free particle is one that is present in a "field-free" space. In quantum mechanics, it means a region of uniform ...
  2. [2]
    Free particle - EPFL Graph Search
    In physics, a free particle is a particle that, in some sense, is not bound by an external force, or equivalently not in a region where its potential energy ...
  3. [3]
    [PDF] The Schrödinger equation - High Energy Physics
    What is V(x) for a free particle? Constant. Smart (easy) choice is V(x)=0. What are the boundary conditions for a free particle? None. So we need to find ...
  4. [4]
    [PDF] 15 Schrodinger Equation - DigitalCommons@USU
    The general solution (15.10) of the free particle Schrödinger equation, being a superposition over plane waves, corresponds to a superposition of momenta and ...
  5. [5]
    [PDF] Lecture 04: Free Particle - Duarte Lab @ UCSD
    Jan 13, 2025 · The free particle propagator encodes a lot of physics. For ... means particle has momentum and energy p = ℏk. E = ℏω. Page 30. Free ...
  6. [6]
    5.2 Newton's First Law – University Physics Volume 1
    Newton's First Law of Motion ... A body at rest remains at rest or, if in motion, remains in motion at constant velocity unless acted on by a net external force.
  7. [7]
    9 Newton's Laws of Dynamics - Feynman Lectures
    Newton wrote down three laws: The First Law was a mere restatement of the Galilean principle of inertia just described.
  8. [8]
    [PDF] Noether's theorem
    For each continuously generated symmetry of the equation of motion, there is an associated conservation law. By a symmetry of the equations of motion, I mean a ...
  9. [9]
    8.3 Conservation of Energy - University Physics Volume 1 | OpenStax
    Sep 19, 2016 · The mechanical energy E of a particle stays constant unless forces outside the system or non-conservative forces do work on it, in which case, ...
  10. [10]
    [PDF] 8.223 IAP 2017 Lecture 5 Full example and conservation of Energy
    So, conservation of energy is built into the E-L equations. As long as there are no time varying forces (e.g. external drivers) or velocity dependent forces ...
  11. [11]
    [PDF] 8.223 IAP 2017 Lecture 4 Mechanics of Lagrangians, and why L = T
    Thus we know that the Lagrangian of a free particle can only depend on the mag- nitude of its velocity, and we know that the result will be motion with ...
  12. [12]
    [PDF] Lagrangian Mechanics
    For instance, Lagrangian of a free particle L = mv2/2 can be transformed as follows. L = mv2. 2. = m(v/ + V). 2. 2. = mv/2. 2. + mv. / · V +.
  13. [13]
    [PDF] Hamiltonian Mechanics
    expressions in round brackets are zero that leads to Hamilton equations (1.6). ... 21) reproduces the well-known solution q = (p/m)t + const for a free particle.
  14. [14]
    6. Hamilton's Equations - Galileo and Einstein
    we have immediately the so-called canonical form of Hamilton's equations of motion: ∂H∂pi=˙qi,∂H∂qi=−˙pi. Evidently going from state space to phase space has ...
  15. [15]
    [PDF] Canonical Quantization and Application to the Quantum Mechanics ...
    The canonical quantization method is simply a recognition that the quantum mechanics of a single particle that was developed from wave mechanics is in fact a ...<|control11|><|separator|>
  16. [16]
    3.4: The Quantum Mechanical Free Particle - Chemistry LibreTexts
    Mar 14, 2019 · The simplest system in quantum mechanics has the potential energy V = 0 everywhere. This is called a free particle since it has no forces ...Example 3 . 4 . 1 : Eigenstates... · Exercise 3 . 4 . 1 · Example 3 . 4 . 2 : Eigenstates...
  17. [17]
    [PDF] 1926-Schrodinger.pdf
    The wave-equation and its application to the hydrogen atom. §5. The intrinsic reason for the appearance of discrete characteristic frequencies. §6. Other ...
  18. [18]
    [PDF] Chapter 4 Schroedinger Equation
    4.1 Free Motion. Eq.(4.15) describes the motion of a free particle. For simplicity, we consider only a one-dimensional motion along the x-axis. Initially, we ...
  19. [19]
    [PDF] Evolution of a free Gaussian wavepacket Consider a free particle ...
    For a free particle H = p2/(2m) and therefore, momentum eigenfunctions are also energy eigenfunctions. So it is the mathematics of Fourier transforms! It is ...Missing: seminal paper
  20. [20]
    2.2: Free Particle- Wave Packets - Physics LibreTexts
    Jan 29, 2022 · Next, the last term in the square brackets of Eq. (30) describes the effect (iii), the wave packet's spread. It may be readily evaluated if the ...
  21. [21]
    [PDF] Lecture 11: Wavepackets and dispersion
    The Gaussian is called a wavepacket because of its Fourier transform: it is a packet of waves with frequencies/wavenumbers clustered around a single value kc ( ...<|control11|><|separator|>
  22. [22]
    The quantum theory of the electron - Journals
    Husain N (2025) Quantum Milestones, 1928: The Dirac Equation Unifies Quantum Mechanics and Special Relativity, Physics, 10.1103/Physics.18.20, 18. Shah R ...
  23. [23]
    [PDF] The Dirac Equation - Centre for Precision Studies in Particle Physics
    Dirac Equation: Plane Wave Solutions. •Now aim to find general plane wave solutions: •Start from Dirac equation (D10): and use. Note in the above equation the ...
  24. [24]
    [PDF] Resolution of the Klein paradox - arXiv
    Klein's paradox [2] results from the conventional solution of the Dirac equation for a potential step of height V that is larger than 2m, where m is the ...
  25. [25]
    [PDF] Helicity, chirality, and the Dirac equation in the non-relativistic limit
    Apr 20, 2018 · However, in the massless limit, the Dirac equation shows that a particle of positive helicity has positive chirality, and vice versa.
  26. [26]
    Discovery of the Positron - American Physical Society
    In August 1932, Carl D. Anderson recorded a positively charged electron (positron) in a cloud chamber, which was later confirmed.