Fact-checked by Grok 2 weeks ago

Bound state

A bound state is a configuration in which two or more particles, atoms, or bodies interact via attractive forces or potentials, confining them to a finite of such that they cannot escape to without external input. In , bound states correspond to trajectories where the total is less than the at , exemplified by orbiting in gravitational fields. In , a is a quantum of particles confined within a , characterized by a normalizable that decays exponentially at large distances and discrete eigenvalues below the threshold, typically with negative relative to the potential at infinity. Bound states differ fundamentally from scattering or free states, where wave functions extend over all space and form a continuous energy spectrum with non-negative energies; in contrast, bound state wave functions are localized, ensuring the particle has a finite probability of being found only within the confining region. Key properties include the quantization of energy levels, which arise from the boundary conditions imposed by the potential, and the absence of degeneracy in one-dimensional systems unless specified otherwise. In , any attractive potential in one dimension supports at least one bound state. Examples of bound states abound in nature and are central to , molecular, and , as well as ; for instance, electrons bound to protons in atoms form stable configurations with quantized energy levels, as seen in the . In a finite square well potential, multiple bound states emerge depending on the well's depth and width, with the number of states increasing as the potential deepens. These states underpin phenomena such as molecular bonding and nuclear stability, where particles like neutrons and protons are trapped by strong forces, and transitions between bound states occur via absorption or emission of like photons.

Classical Bound States

Definition and Characteristics

In , a bound state describes the confined motion of a particle within a finite region of space, occurring when the total energy E of the system is less than the V at infinity, preventing escape to unbounded distances. For typical attractive central potentials where V(r) \to 0 as r \to \infty, this condition simplifies to E < 0, ensuring the particle remains trapped by the potential well. This framework originated in Newtonian mechanics during the 17th century, with foundational examples drawn from celestial mechanics, such as the elliptical orbits of planets around the Sun as articulated in Kepler's laws of planetary motion. Key characteristics of classical bound states include periodic or quasi-periodic trajectories arising from the conservation of total energy and angular momentum, which restrict the particle to closed or nearly closed paths without dissipation./11%3A_Conservative_two-body_Central_Forces/11.07%3A_General_Features_of_the_Orbit_Solutions) In central force problems, these orbits are often elliptical for inverse-square forces like gravity but form rosette patterns—precessing loops—for other potentials, reflecting the deterministic evolution under conserved quantities./11%3A_Conservative_two-body_Central_Forces/11.07%3A_General_Features_of_the_Orbit_Solutions) This contrasts sharply with scattering states, where E > 0 permits trajectories that extend to , representing unbound motion. The dynamics of such bound states are governed by or formulations. For a particle of m in a central potential V(r), the is given by L = \frac{1}{2} m (\dot{r}^2 + r^2 \dot{\theta}^2) - V(r), from which Hamilton's equations derive the , yielding conserved quantities like E = \frac{1}{2} m (\dot{r}^2 + r^2 \dot{\theta}^2) + V(r) and l = m r^2 \dot{\theta}./04%3A_Hamilton%27s_Principle_and_Noether%27s_Theorem/4.09%3A_Example_2-__Lagrangian_Formulation_of_the_Central_Force_Problem)

Examples in Mechanics

One prominent example of a classical bound state is the simple harmonic oscillator, where a particle of m moves in a quadratic potential V(x) = \frac{1}{2} k x^2, with k > 0 being the spring constant. For any finite total energy E > 0, the motion is confined between turning points x = \pm \sqrt{2E/k}, resulting in periodic oscillations. In , the trajectory forms a closed ellipse, parameterized by position x and momentum p = m \dot{x}, illustrating the bounded, periodic nature of the orbit. Another key illustration arises in gravitational systems, such as planetary motion around a central star, governed by the attractive inverse-square potential V(r) = -GMm/r, where G is the , M and m are the masses, and r is the separation. Kepler's first law dictates that bound orbits are ellipses with the star at one focus, ensuring the remains confined without escaping to . For such elliptical orbits, the total energy is negative, specifically E = -GMm/(2a), where a is the semi-major axis, confirming the bound state. In broader central force problems, bound states occur for attractive $1/r potentials when E < 0, yielding closed elliptical orbits analogous to gravitational cases. This contrasts with repulsive $1/r potentials, as in , where positive energies lead to hyperbolic trajectories and unbound scattering rather than confinement. For two-dimensional anisotropic harmonic oscillators, with unequal frequencies \omega_x \neq \omega_y in potentials V(x,y) = \frac{1}{2} m \omega_x^2 x^2 + \frac{1}{2} m \omega_y^2 y^2, the trajectories trace —closed curves that fill the bounded region without escaping, provided the frequencies are commensurate (rational ratio). In classical mechanics, these bound states persist indefinitely, with no mechanism for escape. This setup shares conceptual similarities with the quantum harmonic oscillator but lacks wave-like uncertainty.

Quantum Bound States

Formal Definition

In quantum mechanics, a bound state is defined as an eigenstate of the Hamiltonian operator \hat{H} satisfying the eigenvalue equation \hat{H} \psi = E \psi, where the energy E satisfies E < 0 for potentials V(\mathbf{r}) that vanish at infinity, and the corresponding wave function \psi(\mathbf{r}) is square-integrable over all space, ensuring \int |\psi(\mathbf{r})|^2 d^3\mathbf{r} = 1 and \psi(\mathbf{r}) \to 0 as |\mathbf{r}| \to \infty. This square-integrability reflects the spatial localization of the state, confining the particle to a finite region with probability approaching zero far from the potential well. Bound states are distinguished from scattering states, which correspond to continuum eigenvalues E \geq 0 above the potential threshold, yielding non-normalizable wave functions that extend indefinitely and represent free or asymptotically plane-wave behaviors. The discrete nature of bound state energies arises because only specific eigenvalues below the continuum threshold allow for normalizable solutions to the time-independent Schrödinger equation, ensuring the state's stability against dissociation. The formal framework is encapsulated in the three-dimensional time-independent Schrödinger equation for a single particle of mass m: -\frac{\hbar^2}{2m} \nabla^2 \psi(\mathbf{r}) + V(\mathbf{r}) \psi(\mathbf{r}) = E \psi(\mathbf{r}), where binding requires E < \lim_{|\mathbf{r}| \to \infty} V(\mathbf{r}), typically E < 0 when V(\infty) = 0, preventing the particle from escaping to infinity. This applies primarily to non-relativistic . For multi-particle systems, such as two-body bound states like the , the problem reduces to an effective one-body equation using the reduced mass \mu = \frac{m_1 m_2}{m_1 + m_2} in place of m, separating center-of-mass motion from relative coordinates.

Wave Function Properties

In quantum mechanics, the wave function \psi(\mathbf{r}) of a bound state must be square-integrable over all space to ensure it represents a physically realizable state with finite total probability. This normalizability condition is expressed as \int |\psi(\mathbf{r})|^2 d^3\mathbf{r} = 1, which conserves probability and distinguishes bound states from scattering states, whose wave functions are not normalizable. A key feature of bound state wave functions is their asymptotic behavior at large distances from the confining potential, where the potential V(\mathbf{r}) approaches zero faster than $1/r (short-range case). In this regime, for s-states (l=0), the wave function decays exponentially as \psi(\mathbf{r}) \sim \frac{e^{-\kappa r}}{r}, with \kappa = \sqrt{2m|E|}/\hbar, where E < 0 is the binding energy, m is the particle mass, and \hbar is the reduced Planck's constant. This decay ensures the normalizability and reflects the evanescent nature of the solution to the time-independent Schrödinger equation in regions where the kinetic energy would be imaginary. For long-range potentials like Coulomb, the form differs, lacking the 1/r prefactor in the leading term. In the semiclassical WKB approximation (illustrated here in one dimension for clarity), this exponential tail extends into classically forbidden regions beyond the turning points where E = V(\mathbf{r}), providing a more detailed description of the decay. There, the wave function takes the form \psi(x) \approx \frac{C}{\sqrt{\kappa(x)}} \exp\left( -\int^x \kappa(x') dx' \right), with \kappa(x) = \sqrt{2m(V(x) - E)}/\hbar > 0, ensuring the growing exponential solution is discarded to maintain normalizability. This tail arises from quantum tunneling but diminishes rapidly, confining the particle overall. In three dimensions, the asymptotic also depends on the l, generally involving a prefactor like r^l. Bound state wave functions are spatially localized, meaning the expectation value of \langle \mathbf{r} \rangle is finite, and the probability density |\psi|^2 is concentrated in a finite region around the potential minimum. This localization contrasts with extended states in periodic potentials, such as Bloch waves, which spread indefinitely across the without , leading to delocalized probability. The penetration into forbidden regions via the exponential tail contributes to this confinement, with the probability density peaking in the classically allowed region where E > V(\mathbf{r}).

Key Properties

Discrete Energy Spectrum

In , bound states exhibit a , consisting of isolated eigenvalues lying below the continuous associated with states. This discreteness arises from the requirement that the wave function must satisfy conditions of vanishing at spatial (or exponentially there) to ensure square-integrability and normalizability, which restricts the possible values to a . normalization further enforces this discreteness by excluding continuum solutions that would otherwise be permissible for unbound states. The spacing between these discrete energy levels depends on the form of the confining potential. For the , where V(x) = \frac{1}{2} m \omega^2 x^2, the levels are equally spaced with energies given by E_n = \hbar \omega \left( n + \frac{1}{2} \right), \quad n = 0, 1, 2, \dots yielding a constant separation of \hbar \omega. In contrast, for the Coulomb potential of the , V(r) = -\frac{e^2}{4\pi \epsilon_0 r}, the bound-state energies follow a Rydberg-like E_n = -\frac{13.6 \, \mathrm{eV}}{n^2}, \quad n = 1, 2, 3, \dots where the spacing decreases as n increases, becoming denser near the ionization threshold. As a toy model illustrating this general discreteness, consider the infinite square well potential of width L, where V(x) = 0 for $0 < x < L and infinite elsewhere. The energy eigenvalues are E_n = \frac{n^2 \pi^2 \hbar^2}{2 m L^2}, \quad n = 1, 2, 3, \dots demonstrating quadratic dependence on the quantum number n; even for finite potentials, similar isolated levels emerge below the continuum onset due to the same asymptotic boundary conditions. For smoother potentials, the semiclassical Bohr-Sommerfeld quantization rule provides an approximate method to predict these discrete levels without solving the full , stating that over the classical turning points equals (n + \frac{1}{2}) h: \oint p \, dx = (n + \frac{1}{2}) h, \quad n = 0, 1, 2, \dots where p = \sqrt{2m(E - V(x))}. This approach yields accurate estimates for high-lying states in potentials like the or , bridging classical and quantum descriptions.

Spatial Localization

In , the spatial localization of bound states refers to the confinement of the particle's probability density to a finite region in space, distinguishing them from or states. This localization is a direct consequence of the square-integrable nature of the bound state wave functions, which decay exponentially at large distances from the . A key measure of this localization is the finiteness of the expectation value ⟨r²⟩, defined as ∫ r² |ψ(r)|² d³r, which quantifies the second of the and remains finite for all bound states due to the rapid of |ψ(r)|². The variance, Δr = √(⟨r²⟩ - ⟨r⟩²), provides a standard deviation of this spread, with the typically exhibiting the smallest Δr, rendering it the most localized among the energy eigenstates; for example, in the , the 1s has ⟨r²⟩ = 3 a₀², where a₀ is the , while higher-n states show larger values scaling as n⁴. This confinement has significant implications for physical observables. In true bound states, the ensures stability with infinite lifetimes, as the particle has negligible probability of escaping to . However, in unstable or quasi-bound systems—such as narrow resonances above the potential —spatial localization leads to finite lifetimes via quantum tunneling, where the wave function penetrates the barrier, allowing decay with a rate proportional to the imaginary part of the complex . In stark contrast to free particles, whose delocalized plane-wave descriptions yield uniform probability densities over infinite space and divergent ⟨r²⟩ (due to non-normalizability), bound orbitals are compact and peaked within the attractive potential, enabling phenomena like atomic shells in multi-electron systems. A fundamental aspect of this localization is that bound state wave functions in position space are not eigenstates of the momentum operator, implying inherent spreads in both position (Δx bounded and finite) and momentum (Δp finite but non-zero). The Heisenberg uncertainty principle then requires Δx Δp ≥ ħ/2, with the bounded Δx enforcing a minimal Δp that reflects the superposition of momentum components needed to maintain confinement./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/07%3A_Quantum_Mechanics/7.03%3A_The_Heisenberg_Uncertainty_Principle) This uncertainty product often achieves near-minimal values in ground states, correlating with their tightest spatial localization and discrete energy spacing.

Theorems and Constraints

Non-Degeneracy in One Dimension

In one-dimensional , for a time-independent potential V(x) that is real-valued and supports bound states, each bound energy eigenvalue E is non-degenerate, meaning there is exactly one \psi(x) (up to an overall ) corresponding to that energy. This non-degeneracy follows from the structure of the one-dimensional time-independent : -\frac{\hbar^2}{2m} \frac{d^2 \psi}{dx^2} + V(x) \psi = E \psi, where the eigenfunctions \psi(x) can be chosen real, ensuring they are uniquely determined for each bound state energy. The proof relies on the Sturm-Liouville formulation of the , which guarantees orthogonal eigenfunctions for distinct eigenvalues; for the same eigenvalue, assume two independent real solutions \psi_1 and \psi_2. Their W = \psi_1 \frac{d\psi_2}{dx} - \psi_2 \frac{d\psi_1}{dx} satisfies W' = 0, implying W is constant. For bound states vanishing at \pm \infty, W = 0 everywhere, so \psi_1 and \psi_2 are linearly dependent, contradicting independence and proving non-degeneracy. This property simplifies the solution of one-dimensional quantum problems by eliminating the need to consider multiple states per , in contrast to higher-dimensional systems like the three-dimensional where accidental degeneracies arise due to .

Node Theorem

The node theorem, also known as the oscillation theorem in the context of Sturm-Liouville problems, states that in one-dimensional , the wave function \psi_n(x) of the nth bound state—ordered by increasing E_n—possesses exactly n-1 nodes (points where \psi_n(x) = 0) within the classically allowed between the turning points, excluding the boundaries at infinity where the wave function decays to zero. This holds for real-valued, normalizable eigenfunctions of the time-independent -\frac{\hbar^2}{2m} \frac{d^2\psi}{dx^2} + V(x)\psi = E\psi, assuming a potential V(x) that supports bound states and tends to infinity or a positive constant at |x| \to \infty. A sketch of the proof relies on the separation and comparison theorems from Sturm-Liouville theory. The separation theorem implies that zeros of consecutive eigenfunctions interlace, with no two eigenfunctions sharing a common zero without coinciding entirely. The comparison theorem, applied to potentials with infinite walls at \pm a (approximating the bound state problem as a \to \infty), shows that higher-energy states oscillate more rapidly, introducing an additional node between those of lower states; nodes persist under adiabatic expansion of the well, as adding or removing a node would require \psi = \psi' = 0 at some interior point, leading to the trivial solution \psi \equiv 0, which contradicts normalization. Physically, higher energies allow the wave function to "wiggle" more times in the allowed region before exponential decay in the forbidden regions dominates. This theorem facilitates labeling bound states in solvable potentials, such as the finite square well or , where the (n=1) has no nodes and is nodeless (always positive or negative), while excited states gain nodes sequentially. Nodes signify phase changes in the , corresponding to regions of constructive and destructive in the probability density. In one dimension, non-degeneracy ensures a unique state per nodal count. The node theorem extends to three-dimensional central potentials by reducing the radial to an effective one-dimensional problem for u(r) = r R(r), where R(r) is the radial ; the includes the centrifugal term \frac{\hbar^2 l(l+1)}{2m r^2}, and u(r) satisfies u(0) = 0 with exactly n-1 nodes in (0, \infty) for the nth bound state, yielding n_r = n - l - 1 radial nodes in R(r).

Existence Conditions

Mathematical Requirements

In , bound states are elements of the \mathcal{H} = L^2(\mathbb{R}^d), the space of square-integrable functions where the wave function \psi satisfies \int |\psi|^2 d\mathbf{r} < \infty, ensuring normalizability and physical interpretability as probability densities. This L^2 condition implies that \psi \to 0 as |\mathbf{r}| \to \infty, distinguishing bound states from states that extend to . The operator H = -\frac{\hbar^2}{2m} \nabla^2 + V(\mathbf{r}) must be on a suitable dense in \mathcal{H} to guarantee real eigenvalues and unitary , as required for observables in . ensures the spectrum is real and that bound state energies correspond to isolated eigenvalues. For the time-independent H\psi = E\psi, the potential V(\mathbf{r}) is typically assumed to approach 0 as |\mathbf{r}| \to \infty, with V bounded below and locally integrable, allowing solutions with E < 0 that are confined by the attractive regions of V. Finite or infinite potential wells satisfy these assumptions, provided there exists at least one E < 0 eigenvalue solution. Boundary conditions require \psi and its first derivative \psi' (or the radial derivative in ) to be continuous across finite potential discontinuities to ensure the probability current is well-defined and the wave function remains single-valued. In one dimension, \psi(x) \to 0 and \psi'(x) remains finite as x \to \pm \infty; in three dimensions, the radial wave function u(r) = r \psi(r) satisfies u(0) = 0 and u(r) \to 0 as r \to \infty. Bound states correspond to the point spectrum of H lying below the spectrum, which begins at 0 for potentials vanishing at , ensuring , negative energies separated from the continuous spectrum of free-particle states. This spectral condition, rooted in the decomposition of the spectrum into point and parts, confirms the existence of localized eigenstates.

Stability Considerations

The stability of bound states under perturbations is a central concern in , as small changes to the can shift energies and potentially disrupt localization. In time-independent , the first-order correction to the energy of a bound state is given by \delta E = \langle \psi | \delta V | \psi \rangle, where \psi is the unperturbed and \delta V is the perturbing potential. This shift can unbind the state if it pushes the energy above the threshold, transforming it from a stable bound state to a embedded in the . For instance, in the , an external applied to atomic bound states induces such shifts, leading to if the field strength exceeds a critical value. Quasi-bound states, often arising from these perturbations, exhibit temporary spatial localization but possess finite lifetimes due to tunneling or decay into the . These states are characterized by energies of the form E = E_r - i \Gamma / 2, where E_r is the real part corresponding to the position and \Gamma is the decay width inversely proportional to the lifetime \tau = \hbar / \Gamma. Unlike true bound states with real, negative energies below the , quasi-bound states represent metastable configurations that broaden the otherwise . In time-dependent scenarios, dynamical stability depends on the rate of potential variation; rapid changes can excite bound states to the , while slow evolutions preserve them via the . This theorem states that if the varies sufficiently slowly compared to the inverse energy gaps, the system remains in the instantaneous eigenstate, maintaining bound character. For example, in the adiabatic application of an inducing the , atomic bound states follow the evolving potential without dissociation. The discrete spectrum of bound states renders them particularly susceptible to broadening under such perturbations. In open quantum systems, environmental interactions introduce decoherence, which erodes quantum coherence and can destabilize bound states by facilitating energy dissipation or entanglement with the bath. However, certain dissipationless localized bound states demonstrate robustness, with decoherence rates minimized when the state aligns with dark subspaces decoupled from the environment. Studies of time-dependent potentials further reveal that bound state formation and persistence follow the pulse profile, with stability enhanced for gradual modulations that avoid non-adiabatic transitions.

References

  1. [1]
    Bound State - an overview | ScienceDirect Topics
    A bound state is defined as a quantum state in which a wavefunction decreases rapidly with distance, resulting in a normalizable wavefunction that is ...
  2. [2]
  3. [3]
    [PDF] Quantum Physics I, Lecture Note 12 - MIT OpenCourseWare
    Mar 20, 2016 · Any bound state of a one-dimensional even potential is either even or odd. Proof: The absence of degeneracy implies that the solutions ψ(x) ...
  4. [4]
    [PDF] Quantum Mechanics in One Dimension, Part II - Texas Tech University
    Sep 23, 2021 · Theorem: There is no degeneracy in one-dimensional bound states. Proof: Suppose, there are two TISE solutions corresponding to the same energy:.
  5. [5]
    DOE Explains...Quantum Mechanics - Department of Energy
    A bound state is one where the particles are trapped. One example of a bound state is the electrons, neutrons, and protons that are in an atom. To be “quantized ...
  6. [6]
    [PDF] The Two-Body Problem - UCSB Physics
    a particle has an energy E < 0, its energy is less than the potential energy at infinity, and is thus incapable of escaping to infinity. In this case, we ...
  7. [7]
    Kepler's Laws - MacTutor History of Mathematics
    The greatest achievement of Kepler (1571-1630) was his discovery of the laws of planetary motion. There were such three laws, but here we shall deal only with ...
  8. [8]
    [PDF] Chapter 1 Harmonic Oscillator
    The potential energy of the mass-spring system is. U(x) = kx2/2. (1.1) k. M x. F = − kx. Figure 1.1: Illustration of a mass-spring system. 1. Page 2. CHAPTER 1 ...
  9. [9]
    [PDF] 9. Harmonic Oscillator - MIT OpenCourseWare
    9.1.1 Classical harmonic oscillator and h.o. model. A classical h.o. is described by a potential energy V = 1 kx2. If the system has a finite energy E, the ...
  10. [10]
    [PDF] 4. The Hamiltonian Formalism - DAMTP
    For small θ and small momentum, the pendulum oscillates back and forth, motion which appears as an ellipse in phase space. But for large momentum, the pendulum.
  11. [11]
    13.5 Kepler's Laws of Planetary Motion - University Physics Volume 1
    Sep 19, 2016 · Kepler's first law states that every planet moves along an ellipse, with the Sun located at a focus of the ellipse. An ellipse is defined as ...
  12. [12]
    [PDF] Rutherford scattering and hyperbola orbit - bingweb
    May 27, 2016 · Rutherford was no theorist, but he used his knowledge of central orbits in inverse-square law fields of force to work out the properties of.
  13. [13]
    Lissajous curves and semiclassical theory: The two-dimensional ...
    Mar 1, 2007 · Correspondingly, a Lissajous curve of the commensurate anisotropic harmonic oscillator is reduced to a fundamental segment, for example, the ...
  14. [14]
    [PDF] PY 451 Notes — February 22, 2018 McIntyre Chap. 5 — One ...
    A “bound state” is an eigenstate of the Hamiltonian which is localized in space, i.e., whose wave ... Such solutions are not square-integrable, i.e., R.
  15. [15]
    [PDF] 1 The Schrödinger equation - MIT OpenCourseWare
    Sep 13, 2013 · Since the time-independent Schrödinger equation is a second-order differential equation in x, ... bound states means that the two real ...
  16. [16]
    Schrödinger Wave Equation - Richard Fitzpatrick
    This is the Schrödinger time-independent wave equation. A bound state solution of the previous equation, in which the particle is confined within a finite ...
  17. [17]
    [PDF] Quantum Physics I, Lecture Note 22 - MIT OpenCourseWare
    May 4, 2016 · Our goal here is to show that the two-body quantum mechanical problem of the hydrogen atom can be recast as one in which we have center-of ...
  18. [18]
    [PDF] Schrödinger Equation - JILA
    In the hydrogenic atom, these constants are the reduced mass m = memi/(me + mi), h, and. Ze2. If the nuclear mass mi is large compared to the electronic mass me ...
  19. [19]
    [PDF] Quantum Physics III Chapter 3: Semiclassical Approximation
    Feb 3, 2019 · Similarly, a decaying exponential means our solution decays as we move away from the turning point towards the classically forbidden region. In ...
  20. [20]
    The WKB Approximation - Galileo
    We turn now to the wave function in the classically forbidden region,. p(x)2 ... A bound state will still spill over beyond the classical turning point ...
  21. [21]
    [PDF] Bound States in One Dimension
    This means the electron has a finite chance of being found outside of the classical turning points! Solutions in classically allowed and forbidden step regions.
  22. [22]
    [PDF] EE201/MSE207 Lecture 5 Bound and “scattering” (unbound) states
    Important: Bound states ⇒ discrete energy spectrum (as for infinite QW and oscillator). Scattering states ⇒ continuous energy spectrum (as for free particle).
  23. [23]
    [PDF] Lecture 8: Quantum Harmonic Oscillator - MIT OpenCourseWare
    Mar 5, 2013 · Roughly speaking, there are two sorts of states in quantum mechanics: 1. Bound states: the particle is somewhat localized and cannot escape ...
  24. [24]
    Hydrogen Atom - Richard Fitzpatrick
    $$V(r)$ , the quantized energy levels of a bound-state depend on both $n$ and $l$ (see Sect. 9.3). The fact that the energy levels of a hydrogen atom only depend ...
  25. [25]
    [PDF] 4. The Infinite Square Well - Physics
    We say that the energy is quantized; this is where the “quantum” in quantum mechanics comes from. Here is an energy level diagram (that is, a plot of energy vs.
  26. [26]
    [PDF] Chapter 14 The WKB Method
    The WKB method is useful to obtain approximate eigenergies of bound states and approxi- mate transmission coefficients for tunneling problems.
  27. [27]
    [PDF] Quantization Conditions, 1900–1927 - PhilSci-Archive
    Mar 9, 2020 · Using this procedure, Sommerfeld was able to generalize the circular orbits of Niels Bohr's hydrogen atom (selected through quantization of the ...
  28. [28]
    [PDF] Lecture 5: Stationary States
    Nov 18, 2016 · - Space Localization of Bound and unbound States: As we discussed above the stationary states can have discreet or continues values of ...
  29. [29]
    D.83 Expectation powers of r for hydrogen
    This note derives the expectation values of the powers of $r$ for the hydrogen energy eigenfunctions $\psi_{nlm}$.Missing: finite | Show results with:finite
  30. [30]
    [PDF] 10. Scattering Theory - DAMTP
    Bound states are rather special. In the absence of a potential, a solution which decays exponentially to the left will grow exponentially to the far right.Missing: localized | Show results with:localized
  31. [31]
    2 A Quantum Particle in One Dimension - DAMTP
    As we've seen, these will have a finite number of bound states that decay exponentially, all of which have E < 0 . In what follows, we'll refer to both kinds of ...
  32. [32]
    [quant-ph/0702260] Nodes of Wavefunctions - arXiv
    Feb 27, 2007 · Nodes of Wavefunctions. Authors:M. Moriconi. View a PDF of the paper titled Nodes of Wavefunctions, by M. Moriconi. View PDF.Missing: american | Show results with:american
  33. [33]
    [PDF] Delta Function Potential, Node Theorem, and Simple Harmonic ...
    Apr 1, 2016 · The node theorem states that ψn has n − 1 nodes. We will give an intuitive, non-rigorous explanation of this phenomenon.
  34. [34]
    [PDF] Quantum Mechanics - on Energy Eigenstates - MIT OpenCourseWare
    Feb 28, 2013 · One is that bound states have discrete energies, while unbound states have continuous energies. From this comes the node theorem which works ...
  35. [35]
    [PDF] methods of - modern mathematical physics - 1: functional analysis
    This book is the first of a multivolume series devoted to an exposition of func- tional analysis methods in modern mathematical physics.
  36. [36]
    [PDF] Self-Adjointness of the Atomic Hamiltonian Operator
    Apr 19, 2013 · This class of functions belongs to a Hilbert space H denoted L2 of square integrable functions. It is then seen that for H1 to make any ...
  37. [37]
    [PDF] Proving the existence of bound states for attractive potentials in one ...
    Jul 4, 2021 · This paper proves the existence of bound states for attractive potentials in one and two dimensions using algebraic methods, showing that any ...
  38. [38]
    [PDF] Quantum Physics I, Lecture Note 11 - MIT OpenCourseWare
    Mar 17, 2016 · The ones in the infinite square well are measured with respect to a bottom at zero energy. The ones in the finite square well are measure with ...
  39. [39]
    [PDF] Quantum Bound States - nanoHUB
    It is straightforward to show that the equality of the logarithmic de- rivatives at the matching point reduces to the two conditions we have used for step-.
  40. [40]
    [PDF] methods of - modern mathematical physics - iii: scattering theory
    (iv) In certain very special cases the free dynamics may have bound states (see Section 10). (v) In the Lax-Phillips theory (Section 11) the free dynamics ...
  41. [41]
    Necessary and sufficient conditions for existence of bound states in ...
    Jan 12, 2004 · We also obtain a sufficient condition for the existence of bound states in a central monotonic potential which yield an upper limit on the ...
  42. [42]
    [PDF] 8 Perturbation Theory I: Time Independent Case - DAMTP
    There's a good physical reason why the change in energy comes in at this order: the hydrogen atom is electrically neutral and the unperturbed ground state |1,0, ...
  43. [43]
    Resonances in the Stark Effect of Atomic Systems - Project Euclid
    Stability and Borel Summability. The Stark effect can be described by saying that the bound states of any atomic system turn into resonances as soon as the ...
  44. [44]
    A unified theory of quasibound states | AIP Advances
    Feb 12, 2014 · The theory admits quasibound states with real energies (stationary quasibound states) and quantum resonances within the same framework, and ...
  45. [45]
    Quasibound States in the Continuum in a Two Channel Quantum ...
    Nov 20, 2007 · We report the prediction of quasibound states (resonant states with very long lifetimes) that occur in the eigenvalue continuum of ...
  46. [46]
    [2410.20840] Two-mode Open Quantum Systems: Decoherence ...
    Oct 28, 2024 · In this work, the decoherence dynamics and dissipationless localized bound states of a two-mode open quantum system are investigated.
  47. [47]
    Bound-state formation in time-dependent potentials | Phys. Rev. C
    This study examines bound state formation in a one-dimensional potential using time-dependent pulses, showing that formation follows the pulse shape, not the ...