Fact-checked by Grok 2 weeks ago

Angular momentum operator

In , the is a fundamental vector \hat{\mathbf{L}} that quantifies the rotational dynamics of a quantum system, analogous to the classical \mathbf{L} = \mathbf{r} \times \mathbf{p}. For orbital , it is defined as \hat{\mathbf{L}} = \hat{\mathbf{r}} \times \hat{\mathbf{p}}, where \hat{\mathbf{r}} and \hat{\mathbf{p}} are the and linear momentum s, respectively, with components such as \hat{L}_x = \hat{y}\hat{p}_z - \hat{z}\hat{p}_y. The is Hermitian, ensuring real eigenvalues corresponding to measurable quantities, and its square \hat{L}^2 = \hat{\mathbf{L}} \cdot \hat{\mathbf{L}} represents the total magnitude. The components of the angular momentum operator satisfy the commutation relations [\hat{L}_x, \hat{L}_y] = i[\hbar](/page/H-bar) \hat{L}_z and cyclic permutations, which mirror the of the rotation group SO(3) and dictate the structure of in . Additionally, \hat{L}^2 commutes with each component, [\hat{L}^2, \hat{L}_i] = 0, allowing simultaneous eigenstates for \hat{L}^2 and one component (typically \hat{L}_z). The eigenvalues of \hat{L}^2 are [\hbar](/page/H-bar)^2 l(l+1) for quantum numbers l = 0, 1, 2, \dots, while those of \hat{L}_z are m[\hbar](/page/H-bar) with m = -l, -l+1, \dots, l. These relations arise from the operator definitions and the commutation rules [\hat{x}_i, \hat{p}_j] = i[\hbar](/page/H-bar) \delta_{ij}. Beyond orbital angular momentum, the total angular momentum operator \hat{\mathbf{J}} = \hat{\mathbf{L}} + \hat{\mathbf{S}} incorporates intrinsic \hat{\mathbf{S}}, as in electrons or other particles with , and obeys identical commutation relations [\hat{J}_i, \hat{J}_j] = i[\hbar](/page/H-bar) \epsilon_{ijk} \hat{J}_k. Eigenvalues for \hat{J}^2 follow [\hbar](/page/H-bar)^2 j(j+1) where j can be or , enabling the addition of angular momenta via Clebsch-Gordan coefficients to couple multiple particles or subsystems. In systems with spherical symmetry, such as the , the operators commute with the , forming a that simplify wavefunction separation into radial and angular parts. The angular momentum operators underpin key applications in quantum physics, including atomic spectra, molecular rotations, and particle physics, where they describe selection rules for transitions and symmetry under rotations. Ladder operators \hat{L}_\pm = \hat{L}_x \pm i \hat{L}_y facilitate raising and lowering m values, streamlining computations of matrix elements and spherical harmonics. Their abstract algebraic structure extends to other conserved quantities, like isospin in nuclear physics, highlighting their foundational role in understanding quantum symmetries.

Introduction

Definition and Classical Correspondence

In quantum mechanics, the angular momentum operator is defined as a vector operator \mathbf{L} = (L_x, L_y, L_z), whose components act on wave functions in the position representation. Specifically, the components are given by L_i = -i\hbar (r_j \partial_k - r_k \partial_j), where i, j, k are cyclic permutations of x, y, z, \hbar is the reduced Planck's constant, r_j and r_k are position coordinates, and \partial_j = \partial / \partial r_j are partial derivatives. This differential operator form arises naturally in the wave mechanics formulation, where it generates infinitesimal rotations in the Hilbert space of quantum states. Classically, is the \mathbf{L} = \mathbf{r} \times \mathbf{p}, where \mathbf{r} is the and \mathbf{p} is the , representing the rotational analog of linear momentum in . In the quantum mechanical promotion to operators, the \mathbf{r} remains multiplication by the coordinate, while the \mathbf{p} becomes -i\hbar \nabla, yielding \mathbf{L} = -i\hbar \mathbf{r} \times \nabla. This correspondence extends to the : classical brackets \{A, B\} between dynamical variables are replaced by commutators [A, B]/i\hbar in the , ensuring that the identity for components translates to the corresponding quantum commutation relations. The was introduced in the foundational development of by , , and in their 1926 paper, where it emerged as part of the systematic quantization of multi-degree-of-freedom systems, including derivations of conservation laws for . This early formulation laid the groundwork for treating rotational dynamics quantum mechanically, bridging classical intuitions with the non-commutative operator algebra essential to the theory.

Role in Quantum Mechanics

The angular momentum operator plays a central role in as the generator of infinitesimal rotations, embodying the principle that leads to the conservation of . According to the quantum analog of , if the of a system is invariant under rotations, the components of the angular momentum operator commute with the , ensuring their values remain constant over time. This underpins the analysis of isolated quantum systems, where rotational invariance—common in microscopic interactions—preserves total , facilitating the classification of states and transitions. In atomic physics, the angular momentum operator is essential for interpreting phenomena like the Zeeman effect, in which atomic spectral lines split into multiple components under a weak magnetic field due to the torque exerted on the atom's total angular momentum by the field. Each energy level with total angular momentum quantum number J divides into $2J + 1 sublevels labeled by the magnetic quantum number M, with energy shifts proportional to M, enabling precise measurements of atomic structure and magnetic properties. In , the operator's intrinsic form, known as , connects directly to the , which mandates that particles with half-integer values (fermions) follow antisymmetric wave functions and obey the , while those with integer (bosons) follow symmetric statistics. First rigorously proved by Pauli in 1940, this explains the stability of matter by preventing fermions like electrons from occupying identical states, a cornerstone of and the . Quantum information science exploits the angular momentum operator through spin-1/2 systems, where the two possible projections along an axis (\pm \hbar/2) encode the binary states of a , enabling universal quantum computation via controlled manipulations of pairs. Unlike the , which arises from and governs motion in uniform fields, the angular momentum operator addresses rotational dynamics, proving indispensable in central potentials like the interaction in atoms, where it allows separation of radial and angular variables to yield exact solutions.

Components and Definitions

Orbital Angular Momentum

The orbital angular momentum operator \mathbf{L} for a single particle in quantum mechanics is defined analogously to its classical counterpart, as the cross product of the position operator \mathbf{r} and the linear momentum operator \mathbf{p}, yielding \mathbf{L} = \mathbf{r} \times \mathbf{p}. In the position representation, where \mathbf{p} = -i\hbar \nabla, this becomes the differential operator \mathbf{L} = -i\hbar \mathbf{r} \times \nabla. This form arises from the quantization of classical angular momentum, preserving the structure while incorporating the uncertainty principle through non-commuting operators. In spherical coordinates (r, \theta, \phi), the components of \mathbf{L} simplify significantly due to the rotational symmetry. The z-component, for instance, takes the form L_z = -i\hbar \frac{\partial}{\partial \phi}, which acts on a wavefunction \psi(r, \theta, \phi) by differentiating with respect to the azimuthal angle \phi. The full vector operator \mathbf{L} can be expressed using the gradient in spherical coordinates, \nabla = \hat{e}_r \frac{\partial}{\partial r} + \hat{e}_\theta \frac{1}{r} \frac{\partial}{\partial \theta} + \hat{e}_\phi \frac{1}{r \sin \theta} \frac{\partial}{\partial \phi}, leading to \mathbf{L} = -i\hbar \mathbf{r} \times \nabla. These expressions highlight that \mathbf{L} depends only on angular coordinates and derivatives, independent of the radial position r. The orbital angular momentum operator generates infinitesimal rotations in the particle's configuration space, effectively implementing rotations on scalar wavefunctions via the unitary transformation U(\hat{n}, \delta \theta) \psi(\mathbf{r}) = \psi(\mathbf{r} - \delta \theta \, \hat{n} \times \mathbf{r}) \approx [1 - i (\delta \theta / \hbar) \hat{n} \cdot \mathbf{L}] \psi(\mathbf{r}). This is particularly relevant for orbital motion in central force problems, such as the or any spherically symmetric potential, where the commutes with \mathbf{L}, conserving . In these cases, the separates into radial and angular parts, with the angular portion resembling that of a constrained to a . The eigenfunctions of the orbital operators in the are the Y_{lm}(\theta, \phi), which form a complete for functions on the unit and describe the quantized orbital states for such systems. These functions satisfy the operator equations without specifying the eigenvalue spectrum here, emphasizing their role in representing rotationally invariant solutions.

Spin Angular Momentum

The spin angular momentum operator \mathbf{S} represents an intrinsic property of elementary particles, distinct from orbital angular momentum, and is characterized by the quantum number s, which determines the possible eigenvalues of its magnitude squared operator \mathbf{S}^2 as s(s+1)\hbar^2, yielding a magnitude of \sqrt{s(s+1)}\hbar. For fundamental fermions like the , s = 1/2, while bosons such as photons have s = 1, and this intrinsic spin contributes to the particle's total without relying on spatial coordinates or motion. The concept of electron spin was proposed in 1925 by and to explain the splitting in atomic spectra, attributing it to an internal angular momentum of \hbar/2 for the , independent of its orbital motion. This proposal provided a theoretical framework that aligned with the anomalous and other spectroscopic observations, resolving inconsistencies in earlier models. Experimentally, the idea gained support from the 1922 Stern-Gerlach experiment, which demonstrated the quantization of magnetic moments in silver atoms into two discrete states, later interpreted as evidence for spin-$1/2 projection eigenvalues \pm \hbar/2 along the measurement axis. For spin-$1/2 particles, the components of the spin operator are expressed as S_x = \frac{\hbar}{2} \sigma_x, S_y = \frac{\hbar}{2} \sigma_y, and S_z = \frac{\hbar}{2} \sigma_z, where the \sigma_i (for i = x, y, z) are the 2×2 Hermitian, traceless matrices: \sigma_x = \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}, \quad \sigma_y = \begin{pmatrix} 0 & -i \\ i & 0 \end{pmatrix}, \quad \sigma_z = \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix}. These matrices satisfy \sigma_i^\dagger = \sigma_i, \operatorname{Tr}(\sigma_i) = 0, and \sigma_i^2 = I (the 2×2 identity), ensuring the spin operators are Hermitian and yield real eigenvalues, with the z-component eigenvalues \pm \hbar/2 corresponding to spin-up and spin-down states in the . This representation, introduced by in 1927, forms the basis for describing in non-relativistic of magnetic electrons.

Total Angular Momentum

The total angular momentum operator \mathbf{J} in is defined as the vector sum of the orbital angular momentum operator \mathbf{L} and the angular momentum operator \mathbf{S}, given by \mathbf{J} = \mathbf{L} + \mathbf{S}. This operator acts on the combined orbital-spin of a particle or system, where \mathbf{L} describes the rotational motion associated with the particle's and , while \mathbf{S} accounts for its intrinsic . The components of \mathbf{J} obey the same Lie algebra as those of \mathbf{L} or \mathbf{S} alone, satisfying the commutation relations [J_x, J_y] = i \hbar J_z and cyclic permutations, due to the fact that the orbital and spin operators commute, [\mathbf{L}, \mathbf{S}] = 0. This commutativity arises because \mathbf{L} acts solely on the spatial wave function and \mathbf{S} on the spinor part in the tensor product space. However, in the uncoupled basis of simultaneous eigenstates of \mathbf{L}^2, L_z, \mathbf{S}^2, and S_z, the components of \mathbf{J} mix the basis states, as J^2 does not commute with L_z or S_z individually, necessitating a transition to coupled representations for eigenstates of \mathbf{J}^2 and J_z. In , the total momentum \mathbf{J} plays a central role in the of energy levels, where spin-orbit coupling splits degenerate states according to the value of j, the associated with \mathbf{J}, as seen in the spectrum. Similarly, in , the J characterizes the and of nuclear states, determining the ordering and degeneracy of energy levels in nuclei through interactions that conserve total J. In multi-particle or systems, coupled representations of \mathbf{J} are essential, often constructed via Clebsch-Gordan coefficients to handle the addition of multiple angular momenta.

Algebraic Structure

Operator Definitions in Cartesian Coordinates

In quantum mechanics, the angular momentum operators for orbital angular momentum are represented in the position basis using Cartesian coordinates as differential operators derived from the classical expression \mathbf{L} = \mathbf{r} \times \mathbf{p}, where \mathbf{p} = -i\hbar \nabla. The components take the explicit form L_x = -i\hbar \left( y \frac{\partial}{\partial z} - z \frac{\partial}{\partial y} \right), with cyclic permutations yielding L_y = -i\hbar \left( z \frac{\partial}{\partial x} - x \frac{\partial}{\partial z} \right), \quad L_z = -i\hbar \left( x \frac{\partial}{\partial y} - y \frac{\partial}{\partial x} \right). These expressions act on wave functions \psi(\mathbf{r}) in and correspond to the operator form L_i = \epsilon_{ijk} x_j p_k, where \epsilon_{ijk} is the and summation over repeated indices is implied. The operators L_x, L_y, and L_z are Hermitian, meaning \langle \psi | L_i | \phi \rangle = \langle L_i \psi | \phi \rangle for suitable wave functions \psi and \phi, ensuring real eigenvalues that correspond to measurable quantities. To verify this, consider the inner product \int \psi^* (L_x \phi) \, dV over all space. Substituting the differential form and applying integration by parts twice—first to transfer derivatives from \phi to \psi^*, then integrating by parts again—yields boundary terms that vanish for physically relevant wave functions (e.g., those decaying sufficiently fast at infinity). The result is \int (L_x \psi)^* \phi \, dV, confirming L_x^\dagger = L_x; the proof follows analogously for L_y and L_z. While the above definitions apply specifically to orbital angular momentum in position space, the angular momentum operators more generally—encompassing spin and total angular momentum—are defined abstractly as Hermitian operators satisfying the commutation relations [L_x, L_y] = i\hbar L_z (and cyclic permutations), without dependence on position or momentum coordinates. For spin, these operators act on an internal rather than configuration space, as exemplified by the scaled by \hbar/2 for particles.

Commutation Relations Among Components

The components of the angular momentum operator \mathbf{L} satisfy the canonical commutation relations [L_x, L_y] = i\hbar L_z, [L_y, L_z] = i\hbar L_x, and [L_z, L_x] = i\hbar L_y, which can be compactly written as [L_i, L_j] = i\hbar \epsilon_{ijk} L_k where \epsilon_{ijk} is the Levi-Civita symbol. These relations are derived from the definitions of the orbital components in terms of \mathbf{r} and \mathbf{p} operators, L_x = y p_z - z p_y, L_y = z p_x - x p_z, and L_z = x p_y - y p_x, combined with the fundamental commutation relations [x_i, p_j] = i[\hbar](/page/H-bar) $$delta](/page/Delta)_{ij}, [x_i, x_j] = [p_i, p_j] = 0. To compute, for example, [L_x, L_y], expand the commutator using the bilinearity properties: [L_x, L_y] = [y p_z, z p_x] - [y p_z, x p_z] - [z p_y, z p_x] + [z p_y, x p_z]. Each term is then evaluated via the Leibniz ( for commutators, [AB, C] = A[B, C] + [A, C]B and [A, BC] = B[A, C] + [A, B]C, applied to the monomial products in and operators, with repeated use of the relations to simplify; the cross terms ultimately yield i[\hbar](/page/H-bar) (x p_y - y p_x) = i[\hbar](/page/H-bar) L_z. The cyclic permutations follow analogously. The same commutation algebra [J_i, J_j] = i\hbar \epsilon_{ijk} J_k holds universally for the spin angular momentum operators \mathbf{S}, which are postulated to satisfy these relations without a direct classical position-momentum origin, and for the total angular momentum \mathbf{J} = \mathbf{L} + \mathbf{S}, which inherits the algebra through the vector addition of operators obeying the same structure.

Magnitude and Its Commutators

The squared magnitude of the angular momentum operator is defined as [ \mathbf{L}^2 = L_x^2 + L_y^2 + L_z^2, where $L_x$, $L_y$, and $L_z$ are the Cartesian components of the angular momentum operator.[](https://farside.ph.utexas.edu/teaching/qmech/Quantum/node71.html) This operator corresponds to the square of the total angular momentum vector in [quantum mechanics](/page/Quantum_mechanics).[](https://icourse.club/uploads/files/65fb4c16648d2b93f82fe3271c3381c211f2f532.pdf) A fundamental property of $\mathbf{L}^2$ is that it commutes with each component of the angular momentum operator: [\mathbf{L}^2, L_i] = 0, \quad i = x, y, z. This commutation relation holds due to the algebraic structure of the angular momentum operators.[](https://ocw.mit.edu/courses/8-05-quantum-physics-ii-fall-2013/e765f050b2fd50a9d87d2ae801e9c52f_MIT8_05F13_Chap_09.pdf) To demonstrate it, consider the commutator $[\mathbf{L}^2, L_x]$. The term $[L_x^2, L_x]$ vanishes since any operator commutes with itself. For the remaining terms, [L_y^2, L_x] = L_y [L_y, L_x] + [L_y, L_x] L_y = -i \hbar (L_y L_z + L_z L_y), using the basic commutation relation $[L_x, L_y] = i \hbar L_z$ (and thus $[L_y, L_x] = -i \hbar L_z$). Similarly, [L_z^2, L_x] = L_z [L_z, L_x] + [L_z, L_x] L_z = i \hbar (L_z L_y + L_y L_z), with $[L_z, L_x] = i \hbar L_y$. Adding these contributions yields [L_y^2 + L_z^2, L_x] = -i \hbar (L_y L_z + L_z L_y) + i \hbar (L_z L_y + L_y L_z) = 0, since the expressions in parentheses are identical. The results for $[\mathbf{L}^2, L_y]$ and $[\mathbf{L}^2, L_z]$ follow by cyclic permutation.[](https://physicspages.com/pdf/Quantum%20mechanics/Angular%20momentum%20-%20commutators.pdf)[](https://icourse.club/uploads/files/65fb4c16648d2b93f82fe3271c3381c211f2f532.pdf) The commutation of $\mathbf{L}^2$ with all components implies that $\mathbf{L}^2$ and any single component, such as $L_z$, share a common eigenbasis, permitting the simultaneous measurement of the angular momentum magnitude squared and its projection along a chosen axis.[](https://physicspages.com/pdf/Quantum%20mechanics/Angular%20momentum%20-%20commutators.pdf) This property is central to the algebraic treatment of angular momentum in quantum systems with rotational symmetry.[](https://ocw.mit.edu/courses/8-05-quantum-physics-ii-fall-2013/e765f050b2fd50a9d87d2ae801e9c52f_MIT8_05F13_Chap_09.pdf) The eigenvalues of $\mathbf{L}^2$ take the form $\hbar^2 l(l+1)$ for quantum number $l$, as explored in the quantization of angular momentum. ## Quantization and Eigenstates ### Eigenvalue Spectrum In quantum mechanics, the angular momentum operators $\mathbf{J}$ (encompassing both orbital $\mathbf{L}$ and spin $\mathbf{S}$ contributions) admit simultaneous eigenstates $|j, m\rangle$ for the squared magnitude $J^2$ and the $z$-component $J_z$, due to their commutation relation $[J^2, J_z] = 0$. These eigenstates satisfy the eigenvalue equations J^2 |j, m\rangle = \hbar^2 j(j+1) |j, m\rangle, \quad J_z |j, m\rangle = \hbar m |j, m\rangle, where the quantum number $j$ (often denoted $l$ for pure orbital angular momentum) takes non-negative values $j = 0, \frac{1}{2}, 1, \frac{3}{2}, \dots$, either integer for orbital cases or half-integer for spin contributions.[](https://people.chem.ucsb.edu/metiu/horia/OldFiles/QM2015/Ch13QM.pdf)[](https://ocw.mit.edu/courses/8-05-quantum-physics-ii-fall-2013/e765f050b2fd50a9d87d2ae801e9c52f_MIT8_05F13_Chap_09.pdf) To derive these quantized eigenvalues, consider the operator identity $J^2 = J_z^2 + \frac{1}{2}(J_+ J_- + J_- J_+)$, where $J_\pm = J_x \pm i J_y$ are the raising and lowering combinations derived from the fundamental commutation relations $[J_x, J_y] = i \hbar J_z$ and cyclic permutations.[](https://people.chem.ucsb.edu/metiu/horia/OldFiles/QM2015/Ch13QM.pdf) Assuming a normalized state $|\psi\rangle$ that is an eigenvector of $J_z$ with eigenvalue $\hbar m$, the expectation value becomes $\langle J^2 \rangle = \hbar^2 m^2 + \frac{1}{2} \langle J_+ J_- + J_- J_+ \rangle$. The terms $\langle J_+ J_- \rangle = \|J_- |\psi\rangle\|^2 \geq 0$ and $\langle J_- J_+ \rangle = \|J_+ |\psi\rangle\|^2 \geq 0$ by the positivity of norms, yielding the inequality $\langle J^2 \rangle \geq \hbar^2 m^2$, or equivalently $j(j+1) \geq m^2$ upon identifying the eigenvalue of $J^2$ as $\hbar^2 j(j+1)$.[](https://people.chem.ucsb.edu/metiu/horia/OldFiles/QM2015/Ch13QM.pdf)[](http://physics.gmu.edu/~ellswort/p540/angmom.pdf) This bounds the possible $m$ values such that $|m| \leq j$, with $m$ differing by integers from $-j$ to $+j$. The specific form $j(j+1)$ (rather than $j^2$) arises from the algebraic closure of the representation under the action of $J_\pm$, ensuring a finite-dimensional spectrum.[](https://ocw.mit.edu/courses/8-05-quantum-physics-ii-fall-2013/e765f050b2fd50a9d87d2ae801e9c52f_MIT8_05F13_Chap_09.pdf) For a fixed $j$, the spectrum of $J_z$ exhibits $(2j + 1)$-fold degeneracy in the $m$ quantum number, as each $m$ corresponds to a distinct eigenvalue while sharing the same $J^2$ value; this degeneracy holds universally across systems, such as the [hydrogen atom](/page/Hydrogen_atom) where orbital angular momentum $l$ (integer $j$) labels subshells with $2l + 1$ states independent of the principal [quantum number](/page/Quantum_number) $n$.[](https://people.chem.ucsb.edu/metiu/horia/OldFiles/QM2015/Ch13QM.pdf)[](http://sites.science.oregonstate.edu/~tatej/COURSES/ph426/qmch8.pdf) ### Ladder Operator Method The ladder operator method provides an algebraic framework for determining the eigenvalues and eigenstates of the [angular momentum](/page/Angular_momentum) operators by exploiting the commutation relations to construct [raising](/page/Raising) and lowering operators. These operators systematically connect states within the same multiplet, revealing the [discrete](/page/Discrete) spectrum without solving differential equations.[](https://quantummechanics.ucsd.edu/ph130a/130_notes/node209.html) Define the [raising](/page/Raising) and lowering operators as $ L_+ = L_x + i L_y $ and $ L_- = L_x - i L_y $, respectively. These satisfy the commutation relations $[L_z, L_\pm] = \pm \hbar L_\pm$ and $[L^2, L_\pm] = 0$, ensuring that $L_\pm$ map eigenstates of $L^2$ and $L_z$ to other eigenstates within the same $L^2$ eigenspace but shifted in the $L_z$ eigenvalue. Additionally, the identity $L^2 - L_z^2 = \frac{1}{2} (L_+ L_- + L_- L_+)$ follows from expressing the Cartesian components in terms of $L_\pm$.[](https://www.andrew.cmu.edu/user/suter/AngMomentum.pdf)[](https://quantummechanics.ucsd.edu/ph130a/130_notes/node209.html) Consider a simultaneous eigenstate $|l, m\rangle$ of $L^2$ and $L_z$, satisfying $L^2 |l, m\rangle = [\hbar](/page/H-bar)^2 l(l+1) |l, m\rangle$ and $L_z |l, m\rangle = [\hbar](/page/H-bar) m |l, m\rangle$, where $l$ and $m$ are to be determined. Applying $L_+$ yields $L_z (L_+ |l, m\rangle) = L_+ (L_z + [\hbar](/page/H-bar)) |l, m\rangle = [\hbar](/page/H-bar) (m + 1) (L_+ |l, m\rangle)$, so $L_+ |l, m\rangle$ is an eigenstate of $L_z$ with eigenvalue $[\hbar](/page/H-bar) (m + 1)$. Since $[L^2, L_+] = 0$, it shares the same $L^2$ eigenvalue $[\hbar](/page/H-bar)^2 l(l+1)$. Similarly, $L_- |l, m\rangle$ connects to the state with $m - 1$. The states form a ladder, with $L_+$ raising $m$ by 1 and $L_-$ lowering it by 1.[](https://www.andrew.cmu.edu/user/suter/AngMomentum.pdf) To find the action explicitly, compute the norm $\|L_\pm |l, m\rangle\|^2 = \langle l, m | L_\mp L_\pm | l, m \rangle$. Using the identity for $L^2 - L_z^2$ and the eigenvalues, this yields $\|L_+ |l, m\rangle\|^2 = \hbar^2 [l(l+1) - m(m+1)]$. Thus, the normalized action is $L_+ |l, m\rangle = \hbar \sqrt{l(l+1) - m(m+1)} \, |l, m+1\rangle$, and analogously $L_- |l, m\rangle = \hbar \sqrt{l(l+1) - m(m-1)} \, |l, m-1\rangle$. The matrix elements are $\langle l, m' | L_\pm | l, m \rangle = \hbar \sqrt{l(l+1) - m m'} \, \delta_{m', m \pm 1}$.[](https://quantummechanics.ucsd.edu/ph130a/130_notes/node209.html) The ladder must be finite because the norms must be non-negative. Applying $L_+$ repeatedly increases $m$ until $l(l+1) - m(m+1) = 0$ for some maximum $m_{\max} = l$, where $L_+ |l, l\rangle = 0$. Similarly, lowering reaches $m_{\min} = -l$, with $L_- |l, -l\rangle = 0$. For the ladder to close consistently, the number of states is $2l + 1$, and applying the norm condition at $m = l$ gives $l(l+1) - l(l+1) = 0$, confirming the eigenvalue of $L^2$. Both $l$ and $m$ take values such that $m$ ranges from $-l$ to $l$ in integer steps, with $l = 0, 1/2, 1, 3/2, \dots$. This algebraic procedure establishes the quantized spectrum and the structure of the eigenstates.[](https://www.andrew.cmu.edu/user/suter/AngMomentum.pdf)[](https://quantummechanics.ucsd.edu/ph130a/130_notes/node209.html) ### Spherical Harmonics for Orbital Case In the orbital angular momentum case, the simultaneous eigenfunctions of the operators $ \hat{L}^2 $ and $ \hat{L}_z $ in position space are the [spherical harmonics](/page/Spherical_harmonics) $ Y_l^m(\theta, \phi) $, where $ l $ is a non-negative integer and $ m = -l, -l+1, \dots, l $.[](https://bohr.physics.berkeley.edu/classes/221/1112/notes/orbamsph.pdf) These functions satisfy the eigenvalue equations \hat{L}^2 Y_l^m(\theta, \phi) = \hbar^2 l(l+1) Y_l^m(\theta, \phi), \quad \hat{L}_z Y_l^m(\theta, \phi) = \hbar m Y_l^m(\theta, \phi), providing the basis for expanding angular-dependent wave functions in [quantum mechanics](/page/Quantum_mechanics).[31] The spherical harmonics take the explicit form Y_l^m(\theta, \phi) = (-1)^m \sqrt{\frac{2l+1}{4\pi} \frac{(l-m)!}{(l+m)!}} P_l^m(\cos\theta) , e^{i m \phi} for $ m \geq 0 $, with the associated Legendre functions $ P_l^m(\xi) $ defined as P_l^m(\xi) = (1 - \xi^2)^{m/2} \frac{d^m}{d\xi^m} P_l(\xi), where $ P_l(\xi) $ are the [Legendre polynomials](/page/Legendre_polynomials) and $ \xi = \cos\theta $.[](https://bohr.physics.berkeley.edu/classes/221/1112/notes/orbamsph.pdf) For $ m < 0 $, $ Y_l^m(\theta, \phi) = (-1)^m Y_l^{-m *}(\theta, \phi) $.[](https://farside.ph.utexas.edu/teaching/qm/lectures/node40.html) The associated Legendre functions arise from solving the angular part of the Laplace equation in spherical coordinates, ensuring the separation of variables for the eigenproblem.[](https://bohr.physics.berkeley.edu/classes/221/1112/notes/orbamsph.pdf) The [spherical harmonics](/page/Spherical_harmonics) form a complete [orthonormal basis](/page/Orthonormal_basis) on the unit sphere, satisfying the orthogonality relation \int Y_l^{m *}(\theta, \phi) Y_{l'}^{m'}(\theta, \phi) , d\Omega = \delta_{l l'} \delta_{m m'}, where the integral is over the [solid angle](/page/Solid_angle) $ d\Omega = \sin\theta \, d\theta \, d\phi $ with limits $ \theta $ from 0 to $ \pi $ and $ \phi $ from 0 to $ 2\pi $.[](https://farside.ph.utexas.edu/teaching/qm/lectures/node40.html) This normalization, $ \int |Y_l^m|^2 d\Omega = 1 $, facilitates the expansion of any [square-integrable function](/page/Square-integrable_function) on the sphere.[](https://bohr.physics.berkeley.edu/classes/221/1112/notes/orbamsph.pdf) Under parity transformation (spatial inversion), the spherical harmonics transform as $ Y_l^m(\pi - \theta, \phi + \pi) = (-1)^l Y_l^m(\theta, \phi) $, a property that determines the parity of orbital states and influences selection rules for electric dipole transitions in atomic physics.[](https://bohr.physics.berkeley.edu/classes/221/1112/notes/orbamsph.pdf) ## Symmetry and Rotations ### Generator of Rotations In quantum mechanics, the angular momentum operator $\mathbf{J}$ serves as the generator of rotations, meaning that unitary transformations corresponding to spatial rotations of the system are expressed in terms of exponentials involving $\mathbf{J}$. Specifically, the rotation operator for a rotation by an angle $\theta$ about a unit vector $\hat{\mathbf{n}}$ is given by U(\hat{\mathbf{n}}, \theta) = e^{-i \theta \mathbf{J} \cdot \hat{\mathbf{n}} / \hbar}, where $\hbar$ is the reduced Planck's constant. This operator acts on the Hilbert space of quantum states, implementing the rotation in a manner consistent with the principles of quantum symmetry transformations.[](https://bohr.physics.berkeley.edu/classes/221/1112/notes/spinrot.pdf)[](https://farside.ph.utexas.edu/teaching/qm/lectures/node39.html) For infinitesimal rotations, where $\theta$ is small, the rotation operator expands via the [Taylor series](/page/Taylor_series) to \delta U \approx 1 - i (\theta / \hbar) \mathbf{J} \cdot \hat{\mathbf{n}}, revealing how $\mathbf{J}$ directly generates these transformations. This form arises from the [Lie algebra](/page/Lie_algebra) structure underlying rotations, where the commutation relations of the [angular momentum](/page/Angular_momentum) components, such as $[J_x, J_y] = i \hbar J_z$ and cyclic permutations, ensure the proper closure under infinitesimal changes. The infinitesimal generator property thus links the algebraic structure of $\mathbf{J}$ to the [geometry](/page/Geometry) of the rotation group.[](https://www.physics.rutgers.edu/~steves/Mark_and_Janos/Lectures/Lecture_16_Rotations_and_Angular_Momentum.pdf)[](https://bohr.physics.berkeley.edu/classes/221/1112/notes/spinrot.pdf) In the quantum context, rotations can be distinguished as active or passive. An active rotation physically rotates the [quantum state](/page/Quantum_state) itself, such as applying $U(\hat{\mathbf{n}}, \theta)$ to a wavefunction $\psi(\mathbf{r})$ to yield $\psi'(\mathbf{r}) = U(\hat{\mathbf{n}}, \theta) \psi(\mathbf{r})$, which corresponds to rotating the physical system while keeping the coordinate frame fixed; for example, this transforms the expectation value of [position](/page/Position) as $\langle \mathbf{r} \rangle' = R(\hat{\mathbf{n}}, \theta) \langle \mathbf{r} \rangle$, where $R$ is the classical [rotation matrix](/page/Rotation_matrix). In contrast, a passive rotation reorients the [coordinate system](/page/Coordinate_system) without altering the physical configuration, effectively changing the description of the state in the new frame, such as rotating the axes to observe the same wavefunction from a different perspective. This distinction is crucial for interpreting experiments involving [rotational symmetry](/page/Rotational_symmetry), like those with neutron interferometers.[](https://bohr.physics.berkeley.edu/classes/221/1112/notes/spinrot.pdf) ### SU(2) and SO(3) Representations The angular momentum operators in [quantum mechanics](/page/Quantum_mechanics) satisfy the commutation relations $[J_x, J_y] = i \hbar J_z$ and cyclic permutations, which define the [Lie algebra](/page/Lie_algebra) $\mathfrak{so}(3)$ of the rotation group SO(3). This algebra is isomorphic to $\mathfrak{su}(2)$, the [Lie algebra](/page/Lie_algebra) of the [special unitary group](/page/Special_unitary_group) SU(2), with the isomorphism mapping the basis elements appropriately via the [structure constants](/page/Structure_constants) $\epsilon_{ijk}$.[](https://www.zeuthen.desy.de/~kolanosk/eep06/skripte/lie.pdf)[](https://people.math.harvard.edu/~campbell/qmnotes5.pdf) The representations of this algebra are labeled by a [quantum number](/page/Quantum_number) $j$, which corresponds to the total [angular momentum](/page/Angular_momentum), taking values $j = 0, 1/2, 1, 3/2, \dots$, and for each $j$, the representation space has dimension $2j + 1$. In the context of angular momentum, $j = l$ for orbital angular momentum and $j = s$ for [spin](/page/Spin) angular momentum.[](https://scholar.harvard.edu/files/noahmiller/files/representation_theory_and_quantum_mechanics.pdf)[](https://xiechen.caltech.edu/documents/27822/lecture1113.pdf) The group SO(3) admits only [integer](/page/Integer) $j$ representations, which are single-valued under rotations, making them suitable for describing orbital [angular momentum](/page/Angular_momentum) where wave functions transform unambiguously under spatial rotations. These representations arise naturally from the action of SO(3) on [spherical harmonics](/page/Spherical_harmonics) $Y_{lm}(\theta, \phi)$, with $l$ [integer](/page/Integer) and $m = -l, \dots, l$. In contrast, SU(2), as the double cover of SO(3), includes both [integer](/page/Integer) and [half-integer](/page/Half-integer) $j$ representations, allowing for the description of intrinsic [spin](/page/Spin) [degrees of freedom](/page/Degrees_of_freedom).[](https://www.pas.rochester.edu/~rajeev/phy406/Symmetries8.pdf)[](https://www.math.columbia.edu/~woit/notes10.pdf) Half-integer representations, such as $j = 1/2$ for electrons, are double-valued under SO(3) rotations, meaning a 360° rotation introduces a phase factor of -1 in the state vector, which is physically observable in phenomena like the spin-statistics theorem. These representations are essential for fermionic particles and cannot be realized as true representations of SO(3) but are projective representations thereof. This distinction underpins the separation between orbital (bosonic-like) and spin (fermionic) angular momentum in quantum systems.[](https://scholar.harvard.edu/files/noahmiller/files/representation_theory_and_quantum_mechanics.pdf)[](https://xiechen.caltech.edu/documents/27822/lecture1113.pdf) ### Double Cover and 360° Rotations The [special unitary group](/page/Special_unitary_group) SU(2) provides a double cover of the rotation group SO(3) through a surjective two-to-one [group homomorphism](/page/Group_homomorphism), where each element in SO(3) corresponds to two elements in SU(2), and the [kernel](/page/Kernel) of this [homomorphism](/page/Homomorphism) is [the center](/page/The_Center) {I, -I}.[](https://www.damtp.cam.ac.uk/user/examples/D18S.pdf) This topological structure arises because SU(2) is simply connected while SO(3) is not, leading to distinct behaviors in their representations for quantum [angular momentum](/page/Angular_momentum) operators.[](https://www.cis.upenn.edu/~cis5150/gma-v2-chap9.pdf) In the context of half-integer spin representations, such as the fundamental [spin-1/2](/page/Spin-1/2) case, this double covering manifests physically: a [rotation](/page/Rotation) by 360° around any [axis](/page/Axis) yields the [unitary operator](/page/Unitary_operator) $ U(360^\circ) = -I $, multiplying the [spinor](/page/Spinor) [wave function](/page/Wave_function) by -1, whereas a 720° [rotation](/page/Rotation) is required to recover the identity [operator](/page/Operator).[](https://www.damtp.cam.ac.uk/user/examples/D18S.pdf) This 4π periodicity of [spinors](/page/Spinor) was experimentally confirmed using polarized [neutron](/page/Neutron) [interferometry](/page/Interferometry), where the [interference](/page/Interference) pattern shifted by a [phase](/page/Phase) of π after a 360° [spin](/page/Spin) [rotation](/page/Rotation), consistent with the predicted sign change.[](https://doi.org/10.1016/0375-9601(75)90854-0) The implications of this double cover extend to geometric phases in quantum systems; during adiabatic rotations of the magnetic field for a spin-1/2 particle, a Berry phase of $ -\frac{1}{2} \Omega $ is acquired, where $ \Omega $ is the solid angle subtended by the path on the Bloch sphere, reflecting the topological winding associated with SU(2). This phenomenon underscores the half-integer spin's sensitivity to the full 4π solid angle for a phase of -1. In lower dimensions, analogous topological features appear in anyonic statistics, where braiding phases relate to representations of the braid group, enabling robust encoding in topological quantum computing schemes that leverage non-Abelian anyons for fault-tolerant operations.[](https://link.aps.org/doi/10.1103/RevModPhys.80.1083) ## Physical Implications ### Uncertainty Principle The non-commutativity of the angular momentum components leads to fundamental limits on their simultaneous measurement, analogous to the position-momentum uncertainty principle. Specifically, the commutator $[L_x, L_y] = i \hbar L_z$ implies the inequality $\Delta L_x \Delta L_y \geq \frac{1}{2} |\langle L_z \rangle| \hbar$, where $\Delta L_x$ and $\Delta L_y$ are the standard deviations of the respective operators in a given quantum state, and $\langle L_z \rangle$ is the expectation value of $L_z$.[](https://link.aps.org/doi/10.1103/PhysRev.34.163) This relation holds cyclically for other pairs of components, such as $\Delta L_y \Delta L_z \geq \frac{1}{2} |\langle L_x \rangle| \hbar$ and $\Delta L_z \Delta L_x \geq \frac{1}{2} |\langle L_y \rangle| \hbar$.[](https://chem.libretexts.org/Courses/Pacific_Union_College/Quantum_Chemistry/06%3A_The_Hydrogen_Atom/6.03%3A_The_Three_Components_of_Angular_Momentum_Cannot_be_Measured_Simultaneously_with_Arbitrary_Precision) This uncertainty relation arises from the general Robertson–Schrödinger inequality, which states that for any two [self-adjoint](/page/Self-adjoint) operators $A$ and $B$, the product of their variances satisfies $\Delta A \Delta B \geq \frac{1}{2} |\langle [A, B] \rangle|$, where $[A, B] = AB - BA$ is the [commutator](/page/Commutator).[](https://link.aps.org/doi/10.1103/PhysRev.34.163) Applying this to $A = L_x$ and $B = L_y$, the [commutator](/page/Commutator) yields the angular momentum-specific form, with the right-hand side depending on the expectation value of the third component rather than a universal constant. This demonstrates that precise knowledge of two components is [impossible](/page/Impossible) unless $\langle L_z \rangle = 0$, in which case the minimum uncertainty can approach zero only in specific states.[](https://chem.libretexts.org/Courses/Pacific_Union_College/Quantum_Chemistry/06%3A_The_Hydrogen_Atom/6.03%3A_The_Three_Components_of_Angular_Momentum_Cannot_be_Measured_Simultaneously_with_Arbitrary_Precision) In contrast, the total angular momentum squared $L^2$ commutes with each component, such as $[L^2, L_z] = 0$, allowing simultaneous eigenstates where both $L^2$ and $L_z$ have definite values with no inherent [uncertainty](/page/Uncertainty) between them.[](https://chem.libretexts.org/Courses/Pacific_Union_College/Quantum_Chemistry/06%3A_The_Hydrogen_Atom/6.03%3A_The_Three_Components_of_Angular_Momentum_Cannot_be_Measured_Simultaneously_with_Arbitrary_Precision) However, the components themselves cannot all be sharply defined simultaneously except in the trivial case of zero total [angular momentum](/page/Angular_momentum), where the [quantum number](/page/Quantum_number) $l = 0$, making $L_x = L_y = L_z = 0$ with vanishing variances for all components.[](https://chem.libretexts.org/Courses/Pacific_Union_College/Quantum_Chemistry/06%3A_The_Hydrogen_Atom/6.03%3A_The_Three_Components_of_Angular_Momentum_Cannot_be_Measured_Simultaneously_with_Arbitrary_Precision) ### Conservation in Isolated Systems In quantum mechanics, the conservation of the expectation value of the angular momentum operator in isolated systems follows directly from the Ehrenfest theorem applied to a time-independent operator $\mathbf{L}$. The theorem states that the time derivative of the expectation value is given by \frac{d}{dt} \langle \mathbf{L} \rangle = \frac{i}{\hbar} \langle [H, \mathbf{L}] \rangle, where $H$ is the Hamiltonian and $[ \cdot, \cdot ]$ denotes the commutator. For a rotationally invariant Hamiltonian, which depends only on scalar combinations of coordinates and momenta (such as $r = |\mathbf{r}|$ in central potentials), the components of $\mathbf{L}$ commute with $H$, yielding $[H, \mathbf{L}] = 0$. Consequently, $\frac{d}{dt} \langle \mathbf{L} \rangle = 0$, ensuring that the expectation value of angular momentum remains constant over time.[](https://www.physics.rutgers.edu/~steves/Mark_and_Janos/Lectures/Lecture_9_Ehrenfest_Theorem.pdf)[](https://icourse.club/uploads/files/65fb4c16648d2b93f82fe3271c3381c211f2f532.pdf) This conservation extends to the total [angular momentum](/page/Angular_momentum) $\mathbf{J} = \mathbf{L} + \mathbf{S}$ in isolated [quantum systems](/page/Quantum-Systems) lacking external torques, such as those without [magnetic fields](/page/The_Magnetic_Fields) or asymmetric potentials. In such cases, the [Hamiltonian](/page/Hamiltonian) commutes with $\mathbf{J}$, $[H, \mathbf{J}] = 0$, preserving both the [magnitude](/page/Magnitude) $J^2$ and a chosen component (e.g., $J_z$) of the total angular momentum. A representative example occurs in atomic collisions, where two neutral atoms interact via short-range forces in an otherwise [empty space](/page/The_Empty_Space); the total angular momentum of the colliding pair, including initial orbital contributions from their relative motion, remains conserved throughout the [scattering](/page/Scattering) process, dictating selection rules for final states and angular distributions.[](https://icourse.club/uploads/files/65fb4c16648d2b93f82fe3271c3381c211f2f532.pdf) The underlying principle is the quantum analog of [Noether's theorem](/page/Noether's_theorem), which associates continuous symmetries of the [Hamiltonian](/page/Hamiltonian) with conserved quantities through operator commutators. Rotational invariance implies that the generators of rotations—the angular momentum components—satisfy $[H, J_i] = 0$ for $i = x, y, z$, establishing them as conserved charges without explicit time dependence in the [equations of motion](/page/Equations_of_motion). This framework highlights how spatial isotropy enforces [angular momentum](/page/Angular_momentum) conservation, mirroring classical results but realized via unitary representations of the rotation group.[](https://icourse.club/uploads/files/65fb4c16648d2b93f82fe3271c3381c211f2f532.pdf)[](https://www.feynmanlectures.caltech.edu/III_17.html) ### Coupling of Multiple Angular Momenta In [quantum mechanics](/page/Quantum_mechanics), the coupling of multiple angular momenta arises when combining the angular momentum operators of two or more subsystems, such as electrons or particles, to form a total angular momentum operator $\mathbf{J} = \sum_i \mathbf{J}_i$. For two angular momenta $\mathbf{J}_1$ and $\mathbf{J}_2$ with quantum numbers $j_1$ and $j_2$, the possible values of the total [angular momentum](/page/Angular_momentum) quantum number $j$ range from $|j_1 - j_2|$ to $j_1 + j_2$ in steps of 1, reflecting the decomposition of the [tensor product](/page/Tensor_product) of irreducible representations of the rotation group. This addition rule ensures that the total $m = m_1 + m_2$ is conserved, with each $j$ subspace being degenerate in $m$ from $-j$ to $j$. The coupled basis states, denoted $|j, m; j_1 j_2\rangle$, diagonalize $\mathbf{J}^2$ and $J_z$, and are expressed as linear combinations of the [uncoupled](/page/Uncoupled) product states $|j_1 m_1\rangle |j_2 m_2\rangle$: |j, m; j_1 j_2\rangle = \sum_{m_1, m_2} \langle j_1 m_1 j_2 m_2 | j m \rangle , |j_1 m_1\rangle |j_2 m_2\rangle, where the expansion coefficients $\langle j_1 m_1 j_2 m_2 | j m \rangle$ are the Clebsch-Gordan coefficients, real numbers that vanish unless $m = m_1 + m_2$. These coefficients satisfy [orthogonality](/page/Orthogonality) relations, such as $\sum_{m} \langle j_1 m_1 j_2 m_2 | j m \rangle \langle j' m | j_1 m_1 j_2 m_2 \rangle = \delta_{j j'}$, ensuring the basis transformation is unitary. They can be computed analytically for general $j_1, j_2$ using recursive formulas or explicit expressions involving factorials and binomial coefficients, though tables are commonly used for practical calculations. For low values of $j_1$ and $j_2$, the Clebsch-Gordan coefficients take simple forms that facilitate applications like spin-orbit coupling in atoms, where $\mathbf{J} = \mathbf{L} + \mathbf{S}$ with $j_2 = 1/2$. For instance, when $j_1 = l$ (orbital angular momentum) and $j_2 = 1/2$, the states are: \left| j = l + \frac{1}{2}, m \right\rangle = \sqrt{\frac{l + m + 1/2}{2l + 1}} , |l, m - 1/2\rangle , \left|\frac{1}{2}, \frac{1}{2}\right\rangle + \sqrt{\frac{l - m + 1/2}{2l + 1}} , |l, m + 1/2\rangle , \left|\frac{1}{2}, -\frac{1}{2}\right\rangle, \left| j = l - \frac{1}{2}, m \right\rangle = -\sqrt{\frac{l - m + 1/2}{2l + 1}} , |l, m - 1/2\rangle , \left|\frac{1}{2}, \frac{1}{2}\right\rangle + \sqrt{\frac{l + m + 1/2}{2l + 1}} , |l, m + 1/2\rangle , \left|\frac{1}{2}, -\frac{1}{2}\right\rangle, with coefficients derived from the general [orthogonality](/page/Orthogonality) and [normalization](/page/Normalization) conditions. These explicit values are tabulated for small $j$ up to 5/2 in standard references, enabling straightforward computation of [matrix](/page/Matrix) elements in [perturbation theory](/page/Perturbation_theory) or fine-structure calculations. For systems involving three or more angular momenta, such as multi-electron atoms, [direct coupling](/page/Direct_coupling) becomes cumbersome, necessitating recoupling schemes to transform between different pairing conventions, like $((j_1 j_2) j_{12} j_3) j$ and $(j_1 (j_2 j_3) j_{23}) j$. This recoupling is quantified by the Wigner 6j symbols $\begin{Bmatrix} j_1 & j_2 & j_{12} \\ j_3 & j & j_{23} \end{Bmatrix}$, which appear in the overlap $\langle (j_1 j_2)_{j_{12}} j_3 ; j m | j_1 (j_2 j_3)_{j_{23}} ; j m \rangle = (-1)^{\phi} \sqrt{(2j_{12}+1)(2j_{23}+1)} \begin{Bmatrix} j_1 & j_2 & j_{12} \\ j_3 & j & j_{23} \end{Bmatrix}$, where $\phi$ is a [phase factor](/page/Phase_factor). The 6j symbols, introduced by Wigner and systematized by Racah, satisfy symmetry properties and sum rules that simplify multi-angular momentum calculations. For four or more momenta, 9j symbols extend this framework, relating three possible coupling paths. In [atomic physics](/page/Atomic_physics), recoupling via 6j and 9j symbols is essential for comparing [coupling](/page/Coupling) schemes in multi-electron systems. In LS (Russell-Saunders) [coupling](/page/Coupling), individual orbital angular momenta $\mathbf{l}_i$ couple to total $\mathbf{L} = \sum \mathbf{l}_i$ and spins to $\mathbf{S} = \sum \mathbf{s}_i$, then $\mathbf{J} = \mathbf{L} + \mathbf{S}$, suitable for light atoms where electrostatic interactions [dominate](/page/Dominate) spin-orbit effects. Conversely, jj [coupling](/page/Coupling) pairs each electron's $\mathbf{j}_i = \mathbf{l}_i + \mathbf{s}_i$ before summing to total $\mathbf{J}$, prevailing in heavy atoms with strong spin-orbit [coupling](/page/Coupling); intermediate cases use recoupling coefficients to interpolate between schemes. These transformations, computed using 6j symbols, allow consistent [term symbol](/page/Term_symbol) assignments and energy level predictions across the periodic table. ## Coordinate Representations ### Orbital Angular Momentum in Spherical Coordinates In spherical coordinates $(r, \theta, \phi)$, the orbital angular momentum operator $\mathbf{L} = \mathbf{r} \times \mathbf{p}$ takes the form of differential operators acting on the angular variables $\theta$ and $\phi$, while being independent of the radial coordinate $r$. The components are expressed as follows: L_x = -i\hbar \left( -\sin\phi \frac{\partial}{\partial \theta} - \cot\theta \cos\phi \frac{\partial}{\partial \phi} \right), L_y = -i\hbar \left( \cos\phi \frac{\partial}{\partial \theta} - \cot\theta \sin\phi \frac{\partial}{\partial \phi} \right), L_z = -i\hbar \frac{\partial}{\partial \phi}, and the magnitude squared is L^2 = -\hbar^2 \left[ \frac{1}{\sin\theta} \frac{\partial}{\partial \theta} \left( \sin\theta \frac{\partial}{\partial \theta} \right) + \frac{1}{\sin^2\theta} \frac{\partial^2}{\partial \phi^2} \right]. These operators satisfy the commutation relations $[L_x, L_y] = i\hbar L_z$ and cyclic permutations, generating infinitesimal rotations in the angular domain.[](https://www.asc.ohio-state.edu/jayaprakash.1/828/828ho2.pdf) For a particle in a central potential $V(r)$, the time-independent Schrödinger equation $\left[ -\frac{\hbar^2}{2\mu} \nabla^2 + V(r) \right] \psi = E \psi$ separates in spherical coordinates due to the form of the Laplacian $\nabla^2 = \frac{1}{r^2} \frac{\partial}{\partial r} \left( r^2 \frac{\partial}{\partial r} \right) - \frac{L^2}{r^2 \hbar^2}$. Assuming a separable wave function $\psi(r, \theta, \phi) = R(r) Y(\theta, \phi)$, the angular part yields the eigenvalue equation $L^2 Y = \hbar^2 l(l+1) Y$ with $l = 0, 1, 2, \dots$, while $L_z Y = \hbar m Y$ with $m = -l, \dots, l$. The solutions $Y_{l}^m(\theta, \phi)$ are the spherical harmonics. This separation reduces the problem to a radial equation for $R(r)$ and an angular one solved by the angular momentum operators.[](https://bohr.physics.berkeley.edu/classes/221/notes/orbamsph.pdf) In the semiclassical limit of large quantum numbers, the quantization of orbital [angular momentum](/page/Angular_momentum) finds analogy in rotating macroscopic systems, such as diatomic molecules modeled as [rigid rotors](/page/Rigid_rotor). Here, the Bohr-Sommerfeld quantization condition $\oint p_\phi d\phi = n h$ (with integer $n$) leads to discrete rotational energy levels $E \approx \frac{n^2 \hbar^2}{2I}$ (where $I$ is the [moment of inertia](/page/Moment_of_inertia)), approximating the quantum [rigid rotor](/page/Rigid_rotor) spectrum $E_J = \frac{J(J+1) \hbar^2}{2I}$ for large $J$, as originally applied by Ehrenfest in [1913](/page/1913) using adiabatic invariance.[](https://philsci-archive.pitt.edu/16989/1/Quantum-handbook.pdf) ### Matrix Representations for Spin The matrix representations of the spin angular momentum operators $\mathbf{S}$ are constructed in the Hilbert space of dimension $2s + 1$, where $s$ is the spin quantum number, using the standard basis states $|s, m_s\rangle$ with $m_s = -s, -s+1, \dots, s$. These representations satisfy the commutation relations $[S_x, S_y] = i \hbar S_z$ and cyclic permutations, as derived from the Lie algebra of SU(2).[](https://mpl.mpg.de/fileadmin/user_upload/Marquardt_Division/Teaching/Angular_Momentum.pdf) For the simplest case of spin-$s = 1/2$ (e.g., electrons or other spin-1/2 particles), the operators are proportional to the Pauli matrices $\boldsymbol{\sigma}$, given explicitly as: S_x = \frac{\hbar}{2} \begin{pmatrix} 0 & 1 \ 1 & 0 \end{pmatrix}, \quad S_y = \frac{\hbar}{2} \begin{pmatrix} 0 & -i \ i & 0 \end{pmatrix}, \quad S_z = \frac{\hbar}{2} \begin{pmatrix} 1 & 0 \ 0 & -1 \end{pmatrix}, in the basis where $|1/2, 1/2\rangle = \begin{pmatrix} 1 \\ 0 \end{pmatrix}$ and $|1/2, -1/2\rangle = \begin{pmatrix} 0 \\ 1 \end{pmatrix}$. These matrices were introduced by Pauli to describe the two-valuedness of electron spin and form the fundamental representation of SU(2).[](https://arxiv.org/pdf/2304.03139.pdf) In general, for arbitrary spin $s$, the operators $S_z$ is diagonal in the $|s, m_s\rangle$ basis: \langle s, m_s' | S_z | s, m_s \rangle = m_s \hbar , \delta_{m_s', m_s}, while the raising and lowering operators are $S_+ = S_x + i S_y$ and $S_- = S_x - i S_y$, with matrix elements \langle s, m_s' | S_+ | s, m_s \rangle = \hbar \sqrt{(s - m_s)(s + m_s + 1)} , \delta_{m_s', m_s + 1}, \langle s, m_s' | S_- | s, m_s \rangle = \hbar \sqrt{(s + m_s)(s - m_s + 1)} , \delta_{m_s', m_s - 1}. The transverse components follow as $S_x = (S_+ + S_-)/2$ and $S_y = (S_+ - S_-)/(2i)$, yielding tridiagonal matrices for $S_x$ and $S_y$. These elements ensure the total spin operator satisfies $S^2 |s, m_s\rangle = \hbar^2 s(s+1) |s, m_s\rangle$.[](https://mpl.mpg.de/fileadmin/user_upload/Marquardt_Division/Teaching/Angular_Momentum.pdf) For spin-$s=1$ (e.g., vector bosons like photons in certain contexts), the matrices in the basis $|1,1\rangle, |1,0\rangle, |1,-1\rangle$ are: S_x = \frac{\hbar}{\sqrt{2}} \begin{pmatrix} 0 & 1 & 0 \ 1 & 0 & 1 \ 0 & 1 & 0 \end{pmatrix}, \quad S_y = \frac{\hbar}{\sqrt{2}} \begin{pmatrix} 0 & -i & 0 \ i & 0 & -i \ 0 & i & 0 \end{pmatrix}, \quad S_z = \hbar \begin{pmatrix} 1 & 0 & 0 \ 0 & 0 & 0 \ 0 & 0 & -1 \end{pmatrix}. Higher-spin representations follow the same ladder operator construction, with increasing matrix size and sparsity, and are irreducible under SU(2) transformations. These forms are essential for computing [expectation](/page/Expectation) values, [time evolution](/page/Time_evolution), and entanglement in multi-spin systems.[50]

References

  1. [1]
    [PDF] ANGULAR MOMENTUM - MIT OpenCourseWare
    Dec 16, 2013 · In quantum mechanics the classical vectors lr, lp and Ll become operators. More precisely, they give us triplets of operators: lr → ( ˆ y ...<|control11|><|separator|>
  2. [2]
    Angular Momentum Operators - Richard Fitzpatrick
    We, thus, conclude that Eqs. (527)-(529) are plausible definitions for the quantum mechanical operators which represent the components of angular momentum. Let ...
  3. [3]
    [PDF] Angular Momentum 1 Angular momentum in Quantum Mechanics
    Classically, angular momentum is defined by. L = r × p . Since in QM we have. [xi,pj] = ihδij. it follows that [Li,Lj] 6= 0.
  4. [4]
    [PDF] Quantum Physics I, Lecture Notes 20-21 - MIT OpenCourseWare
    May 3, 2016 · The definition of the angular momentum operator, as you will see, arises from the classical mechanics counterpart. The properties of the ...<|control11|><|separator|>
  5. [5]
    [PDF] Chapter 4
    Commutators and the Correspondence Principle. Formal Connection. Q.M.. Classical Mechanics. Correspondence between. Classical Poisson bracket of functions ...
  6. [6]
    Zur Quantenmechanik. II. | Zeitschrift für Physik A Hadrons and nuclei
    Download PDF · Zeitschrift für Physik. Zur ... Cite this article. Born, M., Heisenberg, W. & Jordan, P. Zur Quantenmechanik. II.. Z. Physik 35, 557–615 (1926).
  7. [7]
    17 Symmetry and Conservation Laws - Feynman Lectures - Caltech
    Symmetry with respect to rotations around the x-, y-, and z-axes implies the conservation of the x-, y-, and z-components of angular momentum. Symmetry with ...
  8. [8]
    Atomic Spectroscopy - Zeeman Effect | NIST
    Oct 3, 2016 · The Zeeman effect for "weak" magnetic fields (the anomalous Zeeman effect) is of special interest because of the importance of Zeeman data in the analysis and ...
  9. [9]
    Resource Letter SS–1: The Spin-Statistics Connection - AIP Publishing
    Jul 1, 2012 · The spin-statistics connection is a fundamental principle of physics that is easy to state but difficult to prove. David Goodstein tells of ...
  10. [10]
    Universal Quantum Computation with Spin- Pairs and Heisenberg ...
    An efficient and intuitive framework for universal quantum computation is presented that uses pairs of spin-1/2 particles to form logical qubits.
  11. [11]
    [PDF] Orbital Angular Momentum and Spherical Harmonics
    Introductory courses on quantum mechanics usually define the orbital angular momentum of a single particle as L = x×p. This formula is borrowed from classical ...
  12. [12]
    Orbital Angular Momentum in Three Dimensions - Galileo
    The Angular Momentum Operators in Spherical Polar Coordinates. The angular momentum operator →L=→r×→p=−iℏ→r×→∇.
  13. [13]
    [PDF] On the quantum mechanics of magnetic electrons
    43 (1927), 601-623. On the quantum mechanics of magnetic electrons. By W. PAULI, Jr. in Hamburg. (Received on 3 May 1927). Translated by D. H. Delphenich. It ...
  14. [14]
    Spinning Electrons and the Structure of Spectra - Nature
    Spinning Electrons and the Structure of Spectra. G. E. UHLENBECK &; S. GOUDSMIT. Nature volume 117, pages 264–265 ...Missing: angular proposal original
  15. [15]
    [PDF] The discovery of the electron spin S.A. Goudsmit
    Today I will talk a little about history. The history of the discovery of the electron spin by George Uhlenbeck and myself. That is tricky; I don't like.Missing: proposal | Show results with:proposal
  16. [16]
    Zur Quantenmechanik des magnetischen Elektrons
    Es wird gezeigt, wie man zu einer Formulierung der Quantenmechanik des magnetischen Elektrons nach der Schrödingerschen Methode der Eigenfunktionen oh.
  17. [17]
    [PDF] chapter 12 total angular momentum
    For a particle with a total angular momentum quantum number j, there is but one eigenvalue of J 2 and the only quantity which can assume different values is mj ...
  18. [18]
    [PDF] Quantum Physics III Chapter 2: Hydrogen Fine Structure
    Apr 2, 2022 · where the total angular momentum J is obtained by adding the orbital angular momentum. L to the spin angular momentum S: J. = L + S . (2.1.15) ...Missing: nuclear | Show results with:nuclear
  19. [19]
    [PDF] Quantum Mechanics - CUNY Graduate Center
    a) [15 points] From the properties of the angular momentum operators (shown in the formulas below) construct explicitly the matrix representations of the ...
  20. [20]
    Nuclear Energy Levels
    The angular momentum quantum number, J, is the integer or half-integer that is the measure of the total angular momentum of the energy state in units of ...
  21. [21]
    [PDF] Quantum Mechanics Made Simple: Lecture Notes - Purdue ...
    B.3 Derivation of Commutation Relations. In general, we can define the angular momentum operator to be. J = i ˆJx + j ˆJy + k ˆJz. (B.3.1). The above is an ...<|control11|><|separator|>
  22. [22]
    General Principles - Richard Fitzpatrick
    The three basic total angular momentum operators, $J_x$ , $J_y$ , and $J_z$ , obey analogous commutation relations to the corresponding orbital and spin ...
  23. [23]
    [PDF] Modern Quantum Mechanics
    Modern Quantum Mechanics is a classic graduate level textbook, covering the main quantum mechanics concepts in a clear, organized, and engaging manner.
  24. [24]
    None
    ### Summary: Derivation of L² Commuting with Lₓ, Lᵧ, L₂
  25. [25]
    [PDF] Chapter 13. Angular Momentum: General Theory
    In this chapter we define angular momentum through the commutation relations between the operators representing its projections on the coordinate axes. These ...
  26. [26]
    [PDF] Angular Momentum Eigenvalues 1 Introduction We will use operator ...
    1 Introduction. We will use operator calculus to find the spectrum of eigenvalues of the angular momentum operators \L2 and \Lz We will show, ...
  27. [27]
    [PDF] Chapter 8 Angular Momentum - Oregon State University
    Feb 26, 2010 · 8.3.2 Quantum Mechanical Angular Momentum. In quantum mechanics, the Cartesian components of the angular momentum operator L= r ! p are ...
  28. [28]
    Angular Momentum Algebra: Raising and Lowering Operators
    Angular momentum is quantized.​​ is an integer. Note that we can easily write the components of angular momentum in terms of the raising and lowering operators. ...
  29. [29]
    [PDF] Derivation of Angular Momentum Rules in Quantum Mechanics
    Mar 23, 2015 · The basic translation to quantum mechan- ics is done by replacing the vector quantities in (1) with their operators; this then defines. the ...
  30. [30]
    [PDF] Orbital Angular Momentum and Spherical Harmonics
    Orbital Angular Momentum Operators in Spherical Coordinates. The standard angular momentum basis is an eigenbasis of the operators (L2,Lz), with certain.
  31. [31]
    Eigenfunctions of Orbital Angular Momentum
    The spherical harmonics are orthogonal functions, and are properly normalized with respect to integration over the entire solid angle: $\displaystyle \int_0 ...
  32. [32]
    [PDF] Rotations in Quantum Mechanics, and Rotations of Spin-1
    Similarly, in quantum mechanics, we will define the angular momentum as the generator of rotations, rather than as r×p, which is not as general. For not ...
  33. [33]
    Rotation Operators - Richard Fitzpatrick
    A rotation operator, R, rotates a system by an angle. Lz is the generator of rotations about the z-axis, and R(0) = 1.
  34. [34]
    [PDF] Rotations and Angular Momentum - Rutgers Physics
    The generators of rotations about the x, y and z axes are, respectively, the x, y and z angular momentum operators. We will now find the eigenstates and ...
  35. [35]
    [PDF] Lie Groups and Quantum Mechanics
    The Lie algebras of SO(3) and SU(2) are isomorphic. This is the chief technical justification for the “electron = spinning ball” analogy. The non-isomorphism.
  36. [36]
    [PDF] Representation theory and quantum mechanics tutorial Spin and the ...
    Aug 3, 2017 · Proposition 1.3.2. The Lie algebra homomorphism π0 : su2 → so3(R) is an isomorphism which sends iσ1 7→ 2rx, iσ2 7→ 2ry ...
  37. [37]
    [PDF] Representation Theory And Quantum Mechanics
    After all, thinking about the previous section, the spin and angular momentum operators are born from representations of SU(2) and. SO(3), respectively.Missing: seminal papers
  38. [38]
    [PDF] 7 Continuous Group - Xie Chen
    In physics, people often mix the notion of SO(3) and SU(2). It happens because in quantum mechanics both projective and nonprojective representations are ...Missing: papers | Show results with:papers
  39. [39]
    [PDF] Lecture 8 Recall that the the rotation group is SO(3) , the set of ...
    1.5. Half-integer spin representations of SU(2) are not representations of SO(3). So why is half-integer spin allowed in nature?
  40. [40]
    [PDF] SU(2) Representations and Their Applications
    SU(2) representations are tractable, with simple construction, and are important in physics as the spin double cover of SO(3) and internal symmetry.
  41. [41]
    [PDF] PQM Supplementary Notes: Spin, topology, SU(2)→SO(3) etc
    SU(2) = { U 2×2 complex matrix : U† = U−1 , detU = 1} . There is a double cover, or two-to-one, surjective homomorphism. Φ : SU(2) → SO(3) , ker Φ = Z2 = {±1}.
  42. [42]
    [PDF] The Quaternions and the Spaces S3, SU(2), SO(3), and RP
    it is easy to show that the map ρ : SU(2) → SO(3) defined such that ρ(Z) = ρZ is a surjective and continuous homomorphism whose kernel is {1,−1}. Since SU(2) ...<|separator|>
  43. [43]
    Non-Abelian anyons and topological quantum computation
    Sep 12, 2008 · Topological quantum computation has emerged as one of the most exciting approaches to constructing a fault-tolerant quantum computer.
  44. [44]
    The Uncertainty Principle | Phys. Rev. - Physical Review Link Manager
    The Uncertainty Principle, HP Robertson Palmer Physical Laboratory, Princeton, NJ, PDF Share, Phys. Rev. 34, 163 – Published 1 July, 1929.
  45. [45]
    6.3: The Three Components of Angular Momentum Cannot be Measured Simultaneously with Arbitrary Precision
    ### Summary: Why Angular Momentum Components Cannot Be Measured Simultaneously with Arbitrary Precision, and Exception for l=0
  46. [46]
    [PDF] Ehrenfest Theorem - Rutgers Physics
    The Ehrenfest Theorem relates the time derivative of the expectoration value of an operator to the expectation value of the commutator of the operator with the ...
  47. [47]
    Quantum simulations of gauge theories with ultracold atoms
    The key idea is angular momentum conservation: in these atomic scattering processes, the total hyperfine angular momentum F tot is conserved. In particular, the ...<|control11|><|separator|>
  48. [48]
    [PDF] Angular momentum in spherical coordinates We wish to write Lx, Ly ...
    Angular momentum in spherical coordinates. We wish to write Lx, Ly, and Lz in terms of spherical coordinates. Recall that the gradient operator is. ∇. = ˆr∂r ...
  49. [49]
    [PDF] Quantization Conditions, 1900–1927 - PhilSci-Archive
    Mar 9, 2020 · We trace the evolution of quantization conditions from Max Planck's intro- duction of a new fundamental constant (h) in his treatment of ...<|control11|><|separator|>
  50. [50]
    [PDF] Lecture 14 Angular momentum operator algebra
    In this lecture we present the theory of angular momentum operator algebra in quantum mechanics. that ˆ J2, ˆ Jz = 0. z − h ˆJz. z .
  51. [51]