Finite potential well
In quantum mechanics, the finite potential well is a one-dimensional model describing a particle confined by a potential that is zero (or constant) within a finite region and rises to a finite positive value outside, allowing the particle's wave function to penetrate and decay exponentially in the classically forbidden regions beyond the barriers.[1] This contrasts with the infinite potential well, where the wave function is strictly zero outside the well due to infinite barriers, leading to discrete energy levels that do not account for tunneling effects.[2]
The time-independent Schrödinger equation is solved separately inside and outside the well, with solutions matched at the boundaries via continuity of the wave function and its derivative.[3] Inside the well, for bound states with energy E less than the barrier height V_0, the wave function is oscillatory, typically expressed as even (cosine) or odd (sine) functions due to the potential's symmetry.[1] Outside, it decays exponentially, reflecting the evanescent nature in forbidden regions. Energy eigenvalues are determined by solving transcendental equations, such as \eta \tan \eta = \xi for even states, where \eta = ka and \xi = \kappa a (with k and \kappa related to E and V_0), resulting in a finite number of bound states that depends on the well's depth and width.[2][3]
This model illustrates key quantum phenomena, including quantum tunneling and the dependence of bound state energies on barrier finiteness, which lowers energies compared to the infinite well and enables applications in semiconductor devices like quantum wells for optoelectronics.[1] For energies above V_0, unbound scattering states form a continuum, allowing transmission and reflection of particles.[2]
One-Dimensional Finite Potential Well
Potential Profile and Setup
The one-dimensional finite potential well models a particle confined between two finite barriers, providing a more realistic approximation to quantum confinement compared to idealized infinite walls. The potential energy V(x) is defined as zero within the well region and a positive constant V_0 outside, specifically V(x) = 0 for -a < x < a and V(x) = V_0 for |x| > a, where V_0 > 0 and $2a is the well width.[3] This creates a rectangular potential profile resembling a "hole" in an otherwise flat landscape at height V_0, allowing the particle to be trapped but with the possibility of tunneling through the barriers.[3]
The system's behavior is governed by the time-independent Schrödinger equation in one dimension:
-\frac{\hbar^2}{2m} \frac{d^2\psi}{dx^2} + V(x)\psi = E\psi,
where \hbar is the reduced Planck's constant, m is the particle mass, \psi(x) is the wave function, and E is the energy eigenvalue.[3] This equation assumes a time-independent, non-relativistic framework for a single particle, focusing on bound states where $0 < E < V_0 to ensure the particle is localized with exponentially decaying probability outside the well.[3]
The tunneling phenomenon in finite potentials was notably applied by George Gamow in 1928 to explain alpha decay.[4] Unlike the infinite potential well, where the wave function vanishes abruptly at the boundaries, the finite well permits penetration into the classically forbidden regions beyond \pm a, resulting in exponentially decaying tails in \psi(x).[5] Consequently, the finite well supports fewer bound states and lower energy levels than the corresponding infinite well for the same width, as the effective confinement is weaker due to this spillover effect.[5]
Solutions to the Schrödinger Equation
The time-independent Schrödinger equation for a particle in the symmetric one-dimensional finite potential well, defined by V(x) = 0 for |x| < a and V(x) = V_0 for |x| > a, is solved piecewise in the three distinct regions.[6]
Inside the well (|x| < a), where the potential is zero, the equation simplifies to the free-particle form:
-\frac{\hbar^2}{2m} \frac{d^2 \psi}{dx^2} = E \psi,
yielding the general oscillatory solution
\psi(x) = A \sin(kx) + B \cos(kx),
with wave number k = \sqrt{2mE}/\hbar.
Outside the well, for x > a, the potential is constant at V_0 > E, so the equation becomes
-\frac{\hbar^2}{2m} \frac{d^2 \psi}{dx^2} + V_0 \psi = E \psi,
with the physically relevant solution being the evanescent (decaying) wave
\psi(x) = C e^{-\kappa x},
where \kappa = \sqrt{2m(V_0 - E)}/\hbar. Similarly, for x < -a,
\psi(x) = D e^{\kappa x},
ensuring exponential decay away from the well on both sides.[6]
Given the even symmetry of the potential, the wave functions naturally separate into even and odd parity solutions. For even parity, A = 0, B \neq 0, and C = D, resulting in \psi(-x) = \psi(x). For odd parity, B = 0, A \neq 0, and C = -D, yielding \psi(-x) = -\psi(x).
The coefficients A, B, C, and D are determined up to normalization, requiring
\int_{-\infty}^{\infty} |\psi(x)|^2 \, dx = 1.
This oscillatory form inside the well reflects classical allowed motion, while the exponential tails outside indicate quantum tunneling into the forbidden region.[6]
Boundary Conditions and Wave Function Matching
In the finite potential well, the time-independent Schrödinger equation yields solutions that must satisfy specific boundary conditions at the interfaces x = ±a, where the potential changes from 0 (inside, |x| < a) to V₀ (outside, |x| > a). The wave function ψ(x) and its first derivative dψ/dx must both be continuous at these points. This requirement stems from the finite nature of the potential, which imposes no infinite barriers or singular forces; discontinuities in ψ(x) would imply infinite kinetic energy or unphysical probability densities, while discontinuities in dψ/dx would suggest infinite accelerations or delta-function potentials, neither of which apply here.[3][1]
For even-parity bound states, the wave function takes the form ψ(x) ∝ cos(kx) inside the well, with k = √(2mE) / ℏ, and ψ(x) ∝ e^{-κ|x|} outside, where κ = √[2m(V₀ - E)] / ℏ > 0 ensures exponential decay in the classically forbidden regions. Applying continuity at x = a, the value matching gives cos(ka) = e^{-κ a} (up to a common normalization constant that scales both regions equally). The derivative continuity at the same boundary yields -k sin(ka) = -κ cos(ka), reflecting the slope alignment between the oscillatory interior and the decaying exterior.[7][1]
Dividing the derivative condition by the value from the wave function continuity eliminates the normalization constant, resulting in the ratio κ / k = tan(ka) for even-parity states. This relation connects the interior wave number k (tied to the energy E via the allowed oscillatory behavior) to the exterior decay constant κ (tied to the binding energy below the potential edge). Similarly, for odd-parity states, where ψ(x) ∝ sin(kx) inside and the same exponential form outside, the value matching at x = a is sin(ka) = e^{-κ a}, and the derivative matching is k cos(ka) = -κ sin(ka), leading to -κ / k = cot(ka). These matching conditions enforce the global smoothness of the wave function across the entire domain.[7][3]
The parameter 1/κ represents the penetration depth, quantifying how far the wave function tunnels into the forbidden regions beyond x = ±a before decaying significantly; larger depths correspond to shallower wells or energies closer to the potential edge, highlighting the evanescent nature of the tails. This penetration is a direct consequence of the continuity requirements, allowing nonzero probability in the barriers despite classical prohibition.[1]
Quantization Condition and Energy Levels
The quantization condition for bound states in the one-dimensional finite potential well arises from matching the wave functions and their derivatives at the boundaries x = ±a, resulting in transcendental equations that determine the discrete energy eigenvalues E_n satisfying 0 < E_n < V_0.[1]
For even-parity states, the condition is ka \tan(ka) = \kappa a, where k = \sqrt{2mE}/\hbar and \kappa = \sqrt{2m(V_0 - E)}/\hbar.[8] For odd-parity states, it is -ka \cot(ka) = \kappa a.[8]
Defining the dimensionless parameters \xi = ka and \eta = \kappa a, the equations simplify to \xi \tan \xi = \eta for even states and -\xi \cot \xi = \eta for odd states, subject to the constraint \xi^2 + \eta^2 = (2m V_0 a^2)/\hbar^2 = r^2, a constant depending on the well parameters.[1]
These equations lack closed-form solutions and are typically solved graphically: plot the semicircle \eta = \sqrt{r^2 - \xi^2} (for $0 < \xi < r) and find its intersections with the periodic branches of \eta = \xi \tan \xi (even) or \eta = -\xi \cot \xi (odd) in the first quadrant.[1] Each intersection corresponds to an allowed \xi_n, from which E_n = (\hbar^2 \xi_n^2)/(2 m a^2).[1]
The number of bound states is finite and depends on the well depth V_0 and width $2a; there is always at least one (even-parity ground state) for any V_0 > 0, with additional states appearing as V_0 or a increases, up to a maximum approximately (r / \pi) + 1/2.[1]
The energy spectrum consists of discrete levels E_n < V_0, alternating in parity, with the ground state E_1 > 0 and successive levels spaced more closely than in the corresponding infinite well.[1]
In the shallow-well limit (small r), only a few bound states exist, and the ground-state energy is small. In the deep-well limit (large r), the spectrum approaches that of the infinite well, with E_n \approx n^2 \pi^2 \hbar^2 / (8 m a^2) and perturbative corrections for finite penetration into the barriers.[1]
Exact solutions require numerical methods to solve the transcendental equations, though approximate analytic expressions are available for weak binding, such as for the ground state E_1 \approx (\hbar^2 / (8 m a^2)) (1 - c / \sqrt{r}) where c is a constant of order unity, valid near the infinite-well limit.[9]
Even and Odd Parity Solutions
In the symmetric one-dimensional finite potential well, where the potential is even, V(x) = V(-x), the bound state solutions to the time-independent Schrödinger equation possess definite parity, allowing classification into even and odd states. The even parity wave functions satisfy ψ_even(-x) = ψ_even(x), while the odd parity ones satisfy ψ_odd(-x) = -ψ_odd(x). These solutions alternate, with the ground state being even and higher excited states alternating in parity. The explicit forms incorporate the wave numbers k = √(2mE)/ℏ inside the well and κ = √[2m(V₀ - E)]/ℏ outside, where E is the bound state energy (0 < E < V₀) determined by the quantization condition.[1]
The normalized even parity wave function takes the form
\psi_\text{even}(x) =
\begin{cases}
A \cos(kx) & |x| < a \\
A \cos(ka) \, e^{-\kappa (|x| - a)} & |x| > a,
\end{cases}
where the normalization constant A ensures ∫_{-∞}^∞ |ψ_even(x)|² dx = 1. Due to symmetry, this condition simplifies to 2A² [∫_0^a cos²(kx) dx + cos²(ka) ∫_a^∞ e^{-2κ(x - a)} dx] = 1, yielding
A = \left[ a + \frac{\sin(2ka)}{2k} + \frac{\cos^2(ka)}{\kappa} \right]^{-1/2}.
The first integral evaluates to (a/2) + sin(2ka)/(4k), and the second to cos²(ka)/(2κ), confirming the expression after doubling for full space. This form maintains continuity of ψ and dψ/dx at x = ±a.[3]
Similarly, the normalized odd parity wave function is
\psi_\text{odd}(x) =
\begin{cases}
A \sin(kx) & |x| < a \\
\operatorname{sgn}(x) \, A \sin(ka) \, e^{-\kappa (|x| - a)} & |x| > a,
\end{cases}
with the normalization constant
A = \left[ a - \frac{\sin(2ka)}{2k} + \frac{\sin^2(ka)}{\kappa} \right]^{-1/2},
derived analogously from ∫_{-∞}^∞ |ψ_odd(x)|² dx = 1, where the inside integral is (a/2) - sin(2ka)/(4k) and the outside contribution is sin²(ka)/(2κ), doubled for symmetry. The sign function ensures odd parity and proper matching at the boundaries.[3]
The probability densities |ψ_even(x)|² and |ψ_odd(x)|² highlight key quantum features. For even states, |ψ_even(x)|² peaks at x = 0 and decays toward the edges inside the well, with exponential tails extending into the forbidden regions (|x| > a), indicating quantum tunneling and an effective well width greater than 2a. Odd states exhibit a node at x = 0, where |ψ_odd(0)|² = 0, with symmetric lobes peaking away from the center and similar evanescent tails outside. This penetration is most pronounced for lower-energy states near E ≈ 0, reducing the effective confinement.[1]
Asymmetric Potential Well
The asymmetric finite potential well features barriers of unequal heights, breaking the parity symmetry of the symmetric case. The potential is defined as
V(x) =
\begin{cases}
V_L & x < 0 \\
0 & 0 < x < a \\
V_R & x > a
\end{cases}
where V_L \neq V_R > 0 and a > 0 is the well width.[10]
For bound states with energy $0 < E < \min(V_L, V_R), the time-independent Schrödinger equation yields oscillatory solutions inside the well and exponentially decaying solutions outside. Inside ($0 < x < a), the wave function is \psi(x) = B \sin(kx) + C \cos(kx), where k = \sqrt{2mE}/\hbar. On the left (x < 0), \psi(x) = A e^{\kappa_L x}, with \kappa_L = \sqrt{2m(V_L - E)}/\hbar; on the right (x > a), \psi(x) = D e^{-\kappa_R x}, with \kappa_R = \sqrt{2m(V_R - E)}/\hbar. These forms ensure the wave function decays in the classically forbidden regions beyond the barriers.[11]
The coefficients A, B, C, D are determined by requiring continuity of \psi(x) and \psi'(x) at the interfaces x = 0 and x = a. This results in two coupled transcendental equations:
\frac{\psi'(0^+)}{\psi(0^+)} = \kappa_L, \quad \frac{\psi'(a^-)}{\psi(a^-)} = -\kappa_R.
Unlike the symmetric well, no closed-form analytic solutions exist for the energy eigenvalues E, as the equations cannot be reduced to simple graphical or algebraic conditions.[10]
Numerical methods are essential to solve for the discrete energy levels. The shooting method is commonly employed: guess an initial E, integrate the Schrödinger equation from the left boundary using the left-region form and initial conditions (e.g., \psi(0^-) = 1, \psi'(0^-) = \kappa_L), continue into the well with the oscillatory solution, and check the logarithmic derivative \psi'(a^-)/\psi(a^-) against -\kappa_R. Adjust E iteratively (e.g., via bisection or secant method) until the mismatch is zero, yielding a bound state energy. The transfer matrix method alternatively propagates the wave function coefficients across regions to enforce matching. These approaches confirm a finite number of bound states, typically fewer than in the symmetric case if one barrier (say V_R < V_L) is sufficiently low, as higher energies may exceed the shallower barrier and become unbound.[12][13]
A simple pseudocode for root-finding the energy levels using bisection on the mismatch function f(E) = \psi'(a^-)/\psi(a^-) + \kappa_R(E) is:
function find_energy(E_min, E_max, tolerance):
while (E_max - E_min > tolerance):
E_mid = (E_min + E_max) / 2
integrate_psi_from_left(E_mid) # Compute ψ and ψ' at x=a from left
f_mid = psi_prime_a / psi_a + kappa_R(E_mid)
if f_mid * f(E_min) > 0:
E_min = E_mid
else:
E_max = E_mid
return (E_min + E_max) / 2
function find_energy(E_min, E_max, tolerance):
while (E_max - E_min > tolerance):
E_mid = (E_min + E_max) / 2
integrate_psi_from_left(E_mid) # Compute ψ and ψ' at x=a from left
f_mid = psi_prime_a / psi_a + kappa_R(E_mid)
if f_mid * f(E_min) > 0:
E_min = E_mid
else:
E_max = E_mid
return (E_min + E_max) / 2
This identifies energies where f(E) = 0, with multiple roots corresponding to different bound states.[12]
The asymmetry shifts the energy levels upward compared to the symmetric well with average barrier height, eliminates even-odd parity classification (wave functions are neither symmetric nor antisymmetric), and localizes probability density more toward the deeper barrier side, where the effective wavelength is shorter due to higher kinetic energy. If one barrier is much lower, the number of bound states decreases, potentially supporting only the ground state.[10]
Such models are applied to realistic semiconductor heterostructures and quantum dots, where interface asymmetries arise from material differences, enabling tailored confinement for optoelectronic devices like LEDs and lasers. For instance, InGaAs/GaAs quantum wells on vicinal surfaces exhibit asymmetric profiles due to strain and incorporation variations, influencing emission properties.[14]
Higher-Dimensional Finite Potential Wells
Rectangular Well in Two Dimensions
The rectangular well in two dimensions extends the one-dimensional finite potential well to a planar rectangular confinement region with finite-height barriers. The potential energy is defined as V(x,y) = 0 inside the rectangle where |x| < a and |y| < b, and V(x,y) = V_0 outside this region, with bound states requiring E < V_0. When a = b, the well is square, exhibiting higher symmetry. This model captures quantum effects in two-dimensional confinements, such as those in semiconductor quantum dots or layered structures.[15]
The time-independent Schrödinger equation in two dimensions,
-\frac{\hbar^2}{2m} \left( \frac{\partial^2 \psi}{\partial x^2} + \frac{\partial^2 \psi}{\partial y^2} \right) + V(x,y) \psi = E \psi,
admits separable solutions of the form \psi(x,y) = \psi_x(x) \psi_y(y), where \psi_x(x) and \psi_y(y) each obey one-dimensional finite well equations with effective depths V_0 and widths $2a and $2b, respectively. Inside the well, the functions take oscillatory forms: for even parity, \psi_x(x) = A \cos(k_x x) with k_x = \sqrt{2m E_x}/\hbar, and similarly for \psi_y(y); odd parity uses sine functions. Outside, they decay exponentially: \psi_x(x) = C e^{-\kappa_x |x|} for |x| > a, where \kappa_x = \sqrt{2m (V_0 - E_x)}/\hbar, and analogously for y. Boundary matching at x = \pm a and y = \pm b yields transcendental equations determining the allowed E_x and E_y.[16]
The total energy is additive, E_{n_x n_y} = E_{n_x}^{(x)} + E_{n_y}^{(y)}, with quantum numbers n_x, n_y = 1, 2, \dots indexing the discrete levels from each direction's quantization condition. The spectrum is denser than in one dimension, as combinations of 1D levels fill the energy range below V_0, and the total number of bound states equals the product of the counts from the individual 1D wells along x and y. For rectangular wells (a \neq b), degeneracies arise when distinct pairs (n_x, n_y) and (m_x, m_y) yield equal energies; in the square case, symmetry enhances degeneracies for states like (n, m) and (m, n).[16]
Quantum tunneling penetrates the finite barriers in both x and y directions, with the wave function extending evanescently into classically forbidden regions. Along the sides (e.g., |x| > a, |y| < b), decay is dominated by the x-component exponential, modulated by y-oscillation; in corner regions ( |x| > a, |y| > b ), the product form results in doubly exponential decay e^{-\kappa_x |x| - \kappa_y |y|}, yielding even smaller probability density there compared to the sides. This reflects enhanced confinement at corners due to barriers in both directions.[16]
Normalization follows from separability: \iint_{-\infty}^{\infty} |\psi(x,y)|^2 \, dx \, dy = 1, which separates into \left( \int_{-\infty}^{\infty} |\psi_x(x)|^2 \, dx \right) \left( \int_{-\infty}^{\infty} |\psi_y(y)|^2 \, dy \right) = 1, with each 1D integral normalized to unity.[16]
Spherical Symmetric Well
The finite spherical potential well is a three-dimensional quantum mechanical model where the potential is spherically symmetric, defined as V(r) = 0 for r < a and V(r) = V_0 for r > a, with V_0 > 0.[17] This setup models bound states for particles confined within a spherical region, analogous to the one-dimensional finite well but adapted to radial coordinates due to the symmetry.[18] The time-independent Schrödinger equation in spherical coordinates separates into radial and angular parts, with the angular portion yielding spherical harmonics Y_{lm}(\theta, \phi).[19] The radial wave function R(r) is obtained by defining u(r) = r R(r), which satisfies the radial equation:
-\frac{\hbar^2}{2m} \frac{d^2 u}{dr^2} + \left[ V(r) + \frac{\hbar^2 l (l+1)}{2m r^2} \right] u(r) = E u(r),
where l is the orbital angular momentum quantum number, and bound states require $0 < E < V_0.[17] For simplicity, the s-wave case (l = 0) eliminates the centrifugal term, reducing the equation to a form similar to the one-dimensional finite well.[18]
For l = 0, inside the well (r < a), the solution is oscillatory and regular at the origin:
u(r) = A \sin(kr), \quad k = \frac{\sqrt{2m E}}{\hbar}.
Outside (r > a), the solution decays exponentially to ensure normalizability:
u(r) = B e^{-\kappa r}, \quad \kappa = \frac{\sqrt{2m (V_0 - E)}}{\hbar}.
Continuity of u(r) and u'(r) at r = a imposes the boundary conditions, leading to the transcendental quantization equation:
k \cot(ka) = -\kappa.
This equation determines the discrete energy levels graphically or numerically, with the number of bound states depending on the well strength parameter \frac{2m V_0 a^2}{\hbar^2}; for example, at least one s-wave bound state exists if this parameter exceeds \left( \frac{\pi}{2} \right)^2, and additional states appear as it increases.[19] The ground state is always an s-wave due to the absence of the centrifugal barrier.[17]
For higher angular momentum l > 0, the centrifugal term \frac{\hbar^2 l (l+1)}{2m r^2} introduces an effective potential V_{\rm eff}(r) = V(r) + \frac{\hbar^2 l (l+1)}{2m r^2}, which acts as a repulsive barrier near the origin, reducing the number of bound states compared to l = 0.[18] Inside the well, the solution involves spherical Bessel functions:
u(r) \propto r j_l(kr),
where j_l is the spherical Bessel function of the first kind, ensuring regularity at r = 0. Outside, the decaying solution uses the spherical Hankel function of the first kind with imaginary argument:
u(r) \propto r h_l^{(1)}(i \kappa r).
The quantization condition generalizes to matching the logarithmic derivatives at r = a:
\frac{j_l'(ka)}{j_l(ka)} = \frac{ \frac{d}{dr} [r h_l^{(1)}(i \kappa r)] / [r h_l^{(1)}(i \kappa r)] }{1} \bigg|_{r=a} \cdot \frac{k}{\kappa},
or equivalently, the ratio of the functions and their derivatives must satisfy continuity, solved numerically for each l.[17] Higher l values require larger V_0 for bound states to exist, as the barrier suppresses penetration near the origin, with typically fewer levels per l than for l = 0.[18]
Annular Well in Spherical Coordinates
The annular well in spherical coordinates models a quantum system confined to a spherical shell between inner radius a and outer radius b > a, with finite potential barriers preventing penetration into the inner core and exterior regions. The potential profile is defined as
V(r) =
\begin{cases}
V_i & 0 \leq r < a \\
0 & a \leq r < b \\
V_o & r \geq b,
\end{cases}
where V_i > 0 and V_o > 0 are the heights of the inner and outer barriers, respectively. This configuration extends the standard spherical well by introducing an inner barrier, creating an annular confinement region suitable for modeling structures like quantum dot rings or atomic electron shells in nuclear physics.
For bound states with energy $0 < E < \min(V_i, V_o), the time-independent Schrödinger equation separates in spherical coordinates, and the radial part for the s-wave (l = 0) reduces to a one-dimensional form for the reduced wave function u(r) = r R(r), where R(r) is the radial component of the full wave function \psi(r, \theta, \phi) = R(r) Y_{0,0}(\theta, \phi). The equation is
-\frac{\hbar^2}{2m} \frac{d^2 u}{dr^2} + V(r) u(r) = E u(r),
with the boundary condition u(0) = 0 to ensure regularity at the origin.
The solutions differ across the three regions due to the piecewise constant potential. In the inner region ($0 \leq r < a), where E < V_i, the solution is evanescent to satisfy u(0) = 0:
u(r) = A \sinh(\kappa_i r),
with \kappa_i = \sqrt{2m(V_i - E)} / \hbar. In the annular region (a \leq r < b), where E > V(r) = 0, the solution is oscillatory:
u(r) = B \sin(kr) + C \cos(kr),
with k = \sqrt{2m E} / \hbar. In the outer region (r \geq b), where E < V_o, the solution decays to ensure normalizability:
u(r) = D e^{-\kappa_o r},
with \kappa_o = \sqrt{2m(V_o - E)} / \hbar. These forms follow directly from solving the radial equation in constant-potential regions, analogous to the one-dimensional finite well but adapted to spherical symmetry with the u(0) = 0 condition.
The coefficients A, B, C, D are determined by requiring continuity of u(r) and its derivative u'(r) at the interfaces r = a and r = b. This yields a system of four equations:
At r = a:
A \sinh(\kappa_i a) = B \sin(ka) + C \cos(ka),
A \kappa_i \cosh(\kappa_i a) = k [B \cos(ka) - C \sin(ka)].
At r = b:
B \sin(kb) + C \cos(kb) = D e^{-\kappa_o b},
k [B \cos(kb) - C \sin(kb)] = - \kappa_o D e^{-\kappa_o b}.
Eliminating the coefficients leads to two coupled transcendental equations in E, which generally require numerical solution to find the discrete bound-state energies. The energy spectrum consists of discrete levels forming bands, with the number of states increasing with the shell width (b - a) and decreasing barrier heights, allowing potentially more bound states than in a simple spherical well of equivalent volume.
In the special case where the inner barrier height diverges (V_i \to \infty), \kappa_i \to \infty, which enforces u(a) = 0 as the boundary condition at the inner edge, reducing the problem to a finite spherical well exterior to an impenetrable core (effectively a hollow sphere with outer finite barrier). This limiting case simplifies the inner matching and highlights how the annular structure modifies the wave function penetration compared to a filled spherical well.
Such potentials arise in models of atomic shells, where the inner barrier represents core repulsion, or in semiconductor quantum dot rings for optoelectronic applications, often necessitating numerical methods like shooting or matrix diagonalization to compute the ground-state wave function and energies due to the lack of closed-form solutions. For the ground state, the wave function is predominantly concentrated in the annular region with exponential tails into the barriers, emphasizing the shell-like confinement.