Fact-checked by Grok 2 weeks ago

Tensor operator

In , a tensor operator is a of scalar and operators, consisting of a set of $2k+1 components T_q^{(k)} (with q = -k, \dots, k) for a given k, that transform under rotations according to the irreducible representations of the rotation group SO(3), mimicking the behavior of or states. These operators satisfy specific commutation relations with the total angular momentum \mathbf{J}, ensuring their transformation properties: [J_z, T_q^{(k)}] = \hbar q T_q^{(k)} and [J_\pm, T_q^{(k)}] = \hbar \sqrt{k(k+1) - q(q \pm 1)} T_{q \pm 1}^{(k)}, where J_\pm = J_x \pm i J_y. Under a rotation \hat{R}, the components transform as \hat{R} T_q^{(k)} \hat{R}^\dagger = \sum_{q'} D_{q' q}^{(k)*}(\hat{R}) T_{q'}^{(k)}, where D^{(k)} are the elements. Tensor operators play a central role in the Wigner-Eckart theorem, which factorizes elements \langle \alpha j m | T_q^{(k)} | \alpha' j' m' \rangle into a reduced matrix element capturing the intrinsic strength and a Clebsch-Gordan coefficient encoding the , simplifying calculations in atomic, nuclear, and . Common examples include the \mathbf{r} as a rank-1 tensor and the electric moment as a rank-2 tensor, enabling the analysis of multipole expansions in electromagnetic interactions.

Fundamentals of Tensor Operators

Definition and General Notion

In , tensor operators serve as a of scalar operators (rank 0) and operators (), extending the framework to higher ranks to describe physical quantities that transform systematically under symmetry operations, particularly rotations, while preserving the underlying commutation relations of the theory. This structure ensures that the algebraic relations between operators remain invariant under group transformations, facilitating the analysis of rotational invariance in . A tensor operator of rank k is defined by its transformation properties under the rotation group SO(3), specifically according to the (2k+1)-dimensional of the group. It comprises $2k+1 components, denoted T_q^{(k)} with q = -k, -k+1, \dots, k, which collectively form a basis transforming among themselves under rotations. These components obey commutation relations with the total operators \mathbf{J} of the form [J_i, T_q^{(k)}] = \sum_{q'} c_{q q'}^{(i,k)} T_{q'}^{(k)}, where the coefficients c depend on q and encode the representation's structure, thereby maintaining the integrity of the under symmetry transformations. The notion of tensor operators was introduced by Eugene P. Wigner in the early 1930s through his application of to , as detailed in his 1931 monograph (original German: Gruppentheorie und ihre Anwendung auf die Quantenmechanik der Atomspektren; English translation 1959) Group Theory and Its Application to the Quantum Mechanics of . This work established tensor operators as essential tools for classifying operators based on their behavior under the rotation group, enabling systematic treatments of atomic and molecular spectra. In contrast to classical tensors, which are multi-index objects representing multilinear maps on finite-dimensional spaces, quantum tensor operators are linear operators acting on the infinite-dimensional of quantum states, with their defining feature being the unitary transformation induced by the group's action on the state space. This quantum-specific adaptation underscores their role in capturing the probabilistic and operator-based nature of quantum observables under symmetries.

Scalar, Vector, and Rank-k Tensor Operators

Tensor operators in quantum mechanics are classified by their rank k, which determines their transformation properties under rotations. A tensor operator of rank k consists of $2k+1 components, labeled by magnetic quantum numbers q = -k, -k+1, \dots, k, and transforms irreducibly under the rotation group SO(3). This classification provides a foundation for understanding how operators behave in angular momentum contexts. Scalar operators, corresponding to rank k=0, are invariant under rotations, meaning their single component T^{(0)} commutes with all components of the total angular momentum operator \mathbf{J}: [ \mathbf{J}, T^{(0)} ] = 0. Examples include a scalar multiple of the identity operator or the radial position operator r = |\mathbf{r}|, which remains unchanged regardless of the coordinate system's orientation. These operators preserve the rotational symmetry of the system and often appear in Hamiltonians for central potentials. Vector operators, of rank k=1, have three components that transform like the position vector \mathbf{r} or momentum \mathbf{p} under rotations. In Cartesian form, the components V_x, V_y, V_z satisfy the commutation relations [J_i, V_j] = i \hbar \epsilon_{ijk} V_k, where \epsilon_{ijk} is the Levi-Civita symbol and J_i are the angular momentum components. A prototypical example is the angular momentum operator itself, \mathbf{J}, which obeys [J_i, J_j] = i \hbar \epsilon_{ijk} J_k. This algebra underscores their role in generating infinitesimal rotations. In spherical basis, the components are V_{\pm 1} = \mp (V_x \pm i V_y)/\sqrt{2} and V_0 = V_z, facilitating ladder operator applications. For higher-rank tensor operators with rank k \geq 2, there are $2k+1 components T_q^{(k)} that transform among themselves under rotations, as originally formalized in the Wigner-Racah framework. The commutation relations with components reflect their "ladder" nature: [J_z, T_q^{(k)}] = \hbar q T_q^{(k)} and [J_{\pm}, T_q^{(k)}] = \hbar \sqrt{k(k+1) - q(q \pm 1)} T_{q \pm 1}^{(k)}. A common example is the rank-2 tensor operator in , which describes the non-spherical charge distribution of nuclei and has five components used to compute electric quadrupole moments. These relations, derived from the theory, ensure consistent with angular momentum states.

Rotations and Transformations

Quantum Rotation Operators

In , spatial rotations are implemented as unitary operators U(R) acting on the of state vectors, where R denotes a in . For a rotation by an angle \phi around a unit axis \mathbf{n}, the operator takes the exponential form U(R) = \exp\left( -i \phi \mathbf{J} \cdot \mathbf{n} / \hbar \right), with \mathbf{J} representing the total . This form arises from the structure of the rotation group SO(3), ensuring that finite rotations are generated by the infinitesimal transformations associated with . The action of the rotation operator on a |\psi\rangle produces the rotated state |\psi'\rangle = U(R) |\psi\rangle. Since U(R) is unitary, satisfying U^\dagger(R) U(R) = I, it preserves the norm of the , \langle \psi' | \psi' \rangle = \langle \psi | \psi \rangle = 1, and thus maintains probabilities in measurements. Similarly, inner products between states are invariant, \langle \psi' | \phi' \rangle = \langle \psi | \phi \rangle, reflecting the of quantum probabilities under rotations. For infinitesimal rotations, where the angle \phi is small, the operator expands as U(R) \approx 1 - i \phi \mathbf{J} \cdot \mathbf{n} / \hbar, identifying the components of \mathbf{J} as the generators of rotations. The angular momentum operators \mathbf{J} satisfy the commutation relations [J_x, J_y] = i \hbar J_z and cyclic permutations, which dictate the algebra of infinitesimal rotations. In the case of orbital angular momentum, the explicit differential operator form is \mathbf{J} = -i \hbar \mathbf{r} \times \nabla, derived from the classical expression \mathbf{L} = \mathbf{r} \times \mathbf{p} upon quantization with \mathbf{p} = -i \hbar \nabla. In the basis of eigenstates |j, m\rangle, where \mathbf{J}^2 |j, m\rangle = \hbar^2 j(j+1) |j, m\rangle and J_z |j, m\rangle = \hbar m |j, m\rangle, the matrix elements of the operator are given by \langle j, m' | U(R) | j, m \rangle = D^{(j)}_{m' m}(R), with D^{(j)}_{m' m}(R) denoting the elements for the of spin j. These elements provide a complete unitary of the group within each .

Transformation Properties under Rotations

In , the transformation properties of operators under rotations are described in the , where a rotated operator T' is given by T' = U(R) T U^\dagger(R), with U(R) denoting the unitary representation of the R. A tensor operator of rank k consists of $2k+1 components T_q^{(k)} (with q = -k, \dots, k) that transform under rotations according to the law U(R) T_q^{(k)} U^\dagger(R) = \sum_{q'=-k}^k D_{q' q}^{(k)}(R) T_{q'}^{(k)}, where D_{q' q}^{(k)}(R) is the element representing the in the of k. This transformation ensures that the components mix among themselves in a manner identical to the or states under the same . To derive this, consider an infinitesimal generated by the total angular momentum \mathbf{J}, where U(\delta R) \approx 1 - i \mathbf{J} \cdot \mathbf{n} \delta \theta for a by \delta \theta around \mathbf{n}. Substituting into the yields the commutation relations [J_\alpha, T_q^{(k)}] = \sum_{q'} (T_{q'}^{(k)} (J_\alpha)_{q' q}^{(k)}), with (J_\alpha)_{q' q}^{(k)} being the matrix elements of J_\alpha in the -k . Finite rotations follow by , confirming the D-matrix form. The set \{ T_q^{(k)} \} forms an irreducible tensor operator if the $2k+1-dimensional space it spans is under rotations and contains no proper subspaces, meaning it furnishes a faithful of the rotation group SO(3). For rank-1 tensor operators, such as vector operators \mathbf{V}, the spherical components relate to the Cartesian basis via V_{\pm 1} = \mp (V_x \pm i V_y)/\sqrt{2} and V_0 = V_z, ensuring the transformation law holds in either basis.

Types of Tensor Operators

Scalar Operators

Scalar operators represent the simplest case of tensor operators in quantum mechanics, classified as rank-zero tensors with a single component denoted as T^{(0)}. These operators are characterized by their rotational invariance, meaning they remain unchanged under any rotation of the system. A defining property is that they commute with all components of the angular momentum operator, satisfying the commutation relation [J_i, T^{(0)}] = 0 for i = x, y, z, where \mathbf{J} is the total angular momentum. Under a unitary rotation operator U(R), a scalar operator transforms as U(R) T^{(0)} U^\dagger(R) = T^{(0)}, confirming its full invariance and lack of directional dependence. This property ensures that expectation values of scalar operators are independent of the orientation of the coordinate system, making them useful for describing isotropic physical quantities. Prominent examples include the Hamiltonian for a particle in a central potential, given by H = \frac{\mathbf{p}^2}{2m} + V(r), where r = |\mathbf{r}| is the scalar radial distance and the kinetic energy term \mathbf{p}^2 is also rotationally invariant. In certain contexts, the parity operator P, which inverts spatial coordinates, behaves as a scalar operator due to its commutation with rotations. Scalar operators can be constructed from higher-rank tensors by taking traces or specific contractions that eliminate angular dependence. For instance, the dot product of two vector operators \mathbf{V} and \mathbf{A}, yielding \mathbf{V} \cdot \mathbf{A}, forms a scalar operator, as the directional components cancel out to produce an invariant quantity.

Vector Operators

Vector operators represent a specific class of rank-1 tensor operators in quantum mechanics, characterized by three components V_q^{(1)} where q = -1, 0, +1 in the spherical basis. These components transform under rotations according to the irreducible representation of the rotation group corresponding to angular momentum quantum number k=1, distinguishing them from scalar operators that remain invariant. The spherical basis components are related to the Cartesian ones by V_0^{(1)} = V_z, V_{\pm 1}^{(1)} = \mp \frac{1}{\sqrt{2}} (V_x \pm i V_y). Common examples of vector operators include the position operator \vec{r}, the momentum operator \vec{p}, and the electric dipole moment operator \vec{d} = -e \vec{r}, where e is the elementary charge. These operators satisfy the defining commutation relations with the total angular momentum operators \vec{J}, which confirm their vectorial transformation properties: [J_i, V_j] = i \hbar \sum_k \epsilon_{ijk} V_k for i,j,k = x,y,z, with \epsilon_{ijk} the Levi-Civita symbol. This relation ensures that the components of \vec{V} rotate as a vector under infinitesimal rotations generated by \vec{J}. The scalar product of two vector operators \vec{V} and \vec{W} yields a rank-0 (scalar) tensor operator, invariant under rotations. In the spherical basis, this product is expressed as \vec{V} \cdot \vec{W} = \sum_{q=-1}^{1} (-1)^q V_q^{(1)} W_{-q}^{(1)}, which follows from the general coupling rules for tensor products of rank-1 operators. This construction highlights how vector operators can be combined to form lower-rank tensors while preserving rotational invariance. In spin systems, the magnetic dipole moment operator is given by \vec{\mu} = -\frac{g \mu_B}{\hbar} \vec{J}, where g is the Landé g-factor, \mu_B = \frac{e \hbar}{2m_e} is the Bohr magneton, and \vec{J} is the total angular momentum operator (or spin \vec{S} for pure spin cases). For an electron's spin, g \approx 2, making this operator a prototypical vector operator that couples to external magnetic fields.

Spherical Tensor Operators

Spherical tensor operators generalize the concept of vector operators to higher ranks and are defined in the as a set of $2k+1 components T_q^{(k)}, where k is the (a non-negative ) and q = -k, -k+1, \dots, k. These operators transform under rotations R according to the rule U(R) T_q^{(k)} U(R)^\dagger = \sum_{q'=-k}^k D_{q' q}^{(k)}(R) T_{q'}^{(k)}, where U(R) is the unitary rotation operator and D^{(k)} are the elements representing the of the rotation group SO(3) for k. This transformation property ensures that the components mix only among themselves, preserving the spherical symmetry inherent to quantum mechanical systems with . In applications such as electromagnetic interactions, spherical tensor operators appear prominently in the of the generated by a charge distribution. The expansion of $1/|\mathbf{r} - \mathbf{r}'| for r > r' yields terms involving gradients of $1/r, where the electric $2^k-pole operator is constructed as a rank-k spherical tensor Q_q^{(k)} = \sum_i e_i r_i^k \sqrt{4\pi/(2k+1)} Y_q^{k*}(\theta_i, \phi_i), with Y_q^k the . These operators facilitate the description of transitions between states, as the matrix elements obey selection rules dictated by the tensor k. A spherical tensor operator of rank k is irreducible if its components cannot be decomposed into a of tensors of lower ranks, meaning the they forms a single of the rotation group. This irreducibility is verified by commutation relations with the total operators \mathbf{J}: [J_z, T_q^{(k)}] = \hbar q T_q^{(k)}, [J_\pm, T_q^{(k)}] = \hbar \sqrt{(k \mp q)(k \pm q + 1)} T_{q \pm 1}^{(k)}, and the double commutator condition \sum_i [J_i, [J_i, T_q^{(k)}]] = \hbar^2 k(k+1) T_q^{(k)}, which mirror those of angular momentum eigenstates. For instance, the \nabla serves as a prototypical example of a rank-1 irreducible spherical tensor, with spherical components \nabla_{\pm 1} = \mp (\nabla_x \pm i \nabla_y)/\sqrt{2} and \nabla_0 = \nabla_z, transforming covariantly under rotations. The components of spherical tensor operators in position space often relate directly to , such that for radial-dependent operators, T_q^{(k)}(\mathbf{r}) \propto r^k Y_q^{(k)}(\hat{r}), enabling their use in expanding potentials or fields with definite (-1)^k. This connection underscores their role in simplifying calculations involving rotational invariance, such as in and molecular .

Construction Methods

Irreducible Tensor Operators

Irreducible tensor operators form a fundamental class in , characterized by their transformation properties under the group. A rank-k irreducible tensor operator consists of $2k+1 components T_q^{(k)}, where q = -k, -k+1, \dots, k, that span the full (2k+1)-dimensional of SO(3) without any invariant subspaces of lower dimension. This means the components mix exclusively among themselves under rotations, ensuring the representation is indecomposable. The criterion for irreducibility is that the set \{T_q^{(k)}\} transforms under a rotation operator U(R) as U(R) T_q^{(k)} U(R)^\dagger = \sum_{q'=-k}^k D_{q' q}^{(k)}(R) \, T_{q'}^{(k)}, where D^{(k)}(R) are the for the -k . Equivalently, the components satisfy specific commutation relations with the total angular momentum operators \mathbf{J}: [J_z, T_q^{(k)}] = \hbar q \, T_q^{(k)}, [J_\pm, T_q^{(k)}] = \hbar \sqrt{(k \mp q)(k \pm q + 1)} \, T_{q \pm 1}^{(k)}, which confirm that the set behaves as a single irreducible multiplet with no separable lower-rank subsets. These relations ensure the absence of invariant subspaces, distinguishing irreducible tensors from reducible ones that contain multiple irreducible components. The primary advantage of irreducible tensor operators lies in their compatibility with the Wigner-Eckart theorem, which greatly simplifies the evaluation of matrix elements in quantum calculations. The theorem states that for states |\alpha j m\rangle and |\alpha' j' m'\rangle (where \alpha denotes additional quantum numbers), the matrix element factors as \langle \alpha \, j \, m | T_q^{(k)} | \alpha' \, j' \, m' \rangle = \frac{\langle j' \, m' \, k \, q | j \, m \rangle}{\sqrt{2j + 1}} \langle \alpha \, j \, \| T^{(k)} \| \alpha' \, j' \rangle, where \langle j' m' k q | j m \rangle is a Clebsch-Gordan coefficient capturing the angular momentum coupling, and \langle \alpha j \| T^{(k)} \| \alpha' j' \rangle is the reduced matrix element depending only on the ranks j, j', k and intrinsic dynamics, independent of magnetic quantum numbers m, m', q. This separation reduces computational complexity, as the geometric factors are tabulated or computable separately, leaving only the reduced matrix element to evaluate directly—often a single integral in atomic or molecular systems. For instance, in multipole expansions or transition amplitudes, this avoids redundant calculations over all m-components. Reducible tensor operators, which do not satisfy the irreducibility criterion, can be decomposed into sums of irreducible ones, facilitating their use in the Wigner-Eckart framework. A classic example is the of two rank-1 operators \mathbf{V} and \mathbf{U}, whose nine Cartesian components V_i U_j (with i,j = x,y,z) decompose into irreducible spherical tensors of ranks k=0,1,2: the k=0 scalar (proportional to the \mathbf{V} \cdot \mathbf{U}), the k=1 antisymmetric part (related to the \mathbf{V} \times \mathbf{U}), and the k=2 symmetric traceless tensor. This decomposition, achieved via Clebsch-Gordan coupling, isolates each irreducible component for independent application of selection rules and matrix element theorems.

Building from Vector Operators

Higher-rank tensor operators can be constructed from the of two operators, denoted as \mathbf{V}^{(1)} and \mathbf{W}^{(1)}, which are rank-1 irreducible tensors. In Cartesian components, this forms a rank-2 tensor T_{ij} = V_i W_j, comprising nine components that transform under rotations as the of two rank-1 representations. This reducible tensor decomposes into irreducible parts according to the Clebsch-Gordan decomposition $1 \otimes 1 = 0 \oplus 1 \oplus 2, yielding a scalar (rank 0, ), an antisymmetric (rank 1), and a symmetric traceless (rank 2). To obtain the irreducible components explicitly in the spherical basis, the rank-2 tensor operator T_q^{(2)} is formed as T_q^{(2)} = \sum_{q_1 q_2} C_{q_1 q_2}^{2 q} V_{q_1}^{(1)} W_{q_2}^{(1)}, where C_{q_1 q_2}^{2 q} are the Clebsch-Gordan coefficients for coupling two angular momenta of j=1. This projection ensures the resulting operator transforms irreducibly under the rank-2 representation. The method generalizes to higher ranks by coupling multiple vector operators sequentially. A representative example is the electric quadrupole moment operator, constructed from the position vector operator \mathbf{r}: Q_{ij} = \sum_\alpha (3 r_{\alpha i} r_{\alpha j} - r_\alpha^2 \delta_{ij}), where the sum is over particles \alpha and the traceless condition Q_{ii} = 0 ensures it is the irreducible rank-2 part. This operator arises from the symmetric traceless projection of the rank-2 tensor product r_i r_j. For scalar operators (rank 0), the dot product of two vector operators provides a simple construction, as in the spin-orbit coupling Hamiltonian term \mathbf{L} \cdot \mathbf{S}, where \mathbf{L} is the orbital vector and \mathbf{S} is the vector. This forms the irreducible scalar from the rank-0 component of the \mathbf{L}^{(1)} \otimes \mathbf{S}^{(1)}.

Using Clebsch–Gordan Coefficients

Clebsch–Gordan coefficients, denoted as \langle j_1 m_1 j_2 m_2 | j m \rangle, quantify the coupling of two angular momenta j_1 and j_2 to yield a total angular momentum j with magnetic quantum number m = m_1 + m_2. These coefficients appear in the expansion of states in the coupled basis |j m\rangle in terms of the uncoupled basis |j_1 m_1\rangle |j_2 m_2\rangle, specifically |j m\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 sum is restricted by m_1 + m_2 = m. The possible values of j range from |j_1 - j_2| to j_1 + j_2 in integer steps, ensuring the decomposition respects the rules of angular momentum addition. In the framework of tensor operators, facilitate the construction of higher-rank irreducible tensor operators from products of lower-rank ones. The of two irreducible tensor operators T^{(k_1)} and U^{(k_2)} decomposes into a of irreducible tensors of ranks k from |k_1 - k_2| to k_1 + k_2. The components of the coupled tensor operator are given by (T^{(k_1)} \times U^{(k_2)})_q^{(k)} = \sum_{q_1 q_2} \langle k_1 q_1 k_2 q_2 | k q \rangle T_{q_1}^{(k_1)} U_{q_2}^{(k_2)}, where the sum runs over q_1 + q_2 = q, and the coefficients ensure the resulting operator transforms irreducibly under rotations. This preserves the spherical and is essential for analyzing matrix elements in systems with multiple angular momenta. For illustration, consider the coupling of two rank-1 tensor operators (k_1 = 1, k_2 = 1), yielding scalar (k=0), vector (k=1), and rank-2 (k=2) components. The Clebsch–Gordan coefficients for this case, following the Condon–Shortley phase convention, are listed in the table below for selected projections (full tables are available in standard references). | k | \langle 1 \, 1 \, 1 \, -1 | k \, 0 \rangle | \langle 1 \, 0 \, 1 \, 0 | k \, 0 \rangle | \langle 1 \, -1 \, 1 \, 1 | k \, 0 \rangle | |-------|---------------------------------------------|--------------------------------------------|---------------------------------------------| | 0 | $1/\sqrt{3} | -1/\sqrt{3} | $1/\sqrt{3} | | 1 | $1/\sqrt{2} | $0 | -1/\sqrt{2} | | 2 | $1/\sqrt{6} | $2/\sqrt{6} | $1/\sqrt{6} | These values demonstrate the antisymmetric scalar combination and the symmetric traceless for k=2. can be computed using recursion relations derived from the commutation properties of ladder operators J_\pm = J_x \pm i J_y. Applying J_\pm to the coupled state |j m\rangle yields relations such as \sqrt{(j \mp m)(j \pm m + 1)} \langle j_1 m_1 j_2 m_2 | j \, m \pm 1 \rangle = \sqrt{(j_1 \mp m_1)(j_1 \pm m_1 + 1)} \langle j_1 \, m_1 \pm 1 \, j_2 m_2 | j \, m \pm 1 \rangle + \sqrt{(j_2 \mp m_2)(j_2 \pm m_2 + 1)} \langle j_1 m_1 j_2 \, m_2 \pm 1 | j \, m \pm 1 \rangle, allowing iterative determination starting from the highest-weight state where m = j_1 + j_2. These recursions ensure consistency with the algebra of the rotation group SU(2). A key symmetry of the coefficients is their orthogonality, which for fixed j reads \sum_m \langle j_1 m_1 j_2 m - m_1 | j m \rangle \langle j_1 m_1' j_2 m - m_1' | j m \rangle = \delta_{m_1 m_1'}, reflecting the unitarity of the between bases and enabling efficient numerical computations or proofs of . This property, along with phase conventions, standardizes their use across quantum mechanical calculations.

Angular Momentum and Spherical Harmonics

Orbital Angular Momentum Operators

The orbital angular momentum operator \mathbf{L} in quantum mechanics is defined as the cross product \mathbf{L} = \mathbf{r} \times \mathbf{p}, where \mathbf{r} is the position operator and \mathbf{p} is the linear momentum operator, analogous to its classical counterpart. This operator represents the rotational motion of a particle relative to a fixed origin. In the position representation, the Cartesian components take the differential form: L_x = -i \hbar \left( y \frac{\partial}{\partial z} - z \frac{\partial}{\partial y} \right), \quad 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 arise from substituting \mathbf{p} = -i \hbar \nabla into the component definitions L_x = y p_z - z p_y, and cyclic permutations. The components of \mathbf{L} satisfy the commutation relations of the of the rotation group SO(3): [L_x, L_y] = i \hbar L_z, \quad [L_y, L_z] = i \hbar L_x, \quad [L_z, L_x] = i \hbar L_y, along with [L^2, L_i] = 0 for i = x, y, z, where L^2 = L_x^2 + L_y^2 + L_z^2. These relations ensure that \mathbf{L} behaves as an irreducible vector operator under rotations. The simultaneous eigenstates of L^2 and L_z, denoted |l, m \rangle, have eigenvalues L^2 |l, m \rangle = \hbar^2 l (l + 1) |l, m \rangle and L_z |l, m \rangle = \hbar m |l, m \rangle, with the orbital quantum number l = 0, 1, 2, \dots and the m = -l, -l+1, \dots, l. The orbital angular momentum operators generate infinitesimal rotations in the orbital degrees of freedom, with the unitary rotation operator given by U(R) = \exp\left( -i \boldsymbol{\theta} \cdot \mathbf{L} / \hbar \right) for a rotation by angle |\boldsymbol{\theta}| around the axis \hat{\boldsymbol{\theta}}. Under this transformation, the position operator satisfies U(R) \mathbf{r} U(R)^\dagger = R \mathbf{r}, where R is the corresponding rotation matrix, confirming \mathbf{L} as the generator of spatial rotations. In the hydrogen atom, the quantum number l labels the subshells, such as s (l=0), p (l=1), d (l=2), and f (l=3), which determine the angular dependence of the electron wavefunctions and the orbital shapes.

Coupling with Spherical Harmonics

Spherical harmonics Y_l^m(\theta, \phi) serve as the angular basis functions in quantum mechanics, normalized such that \int |Y_l^m|^2 \, d\Omega = 1, where the integral is over the solid angle d\Omega = \sin\theta \, d\theta \, d\phi. These functions transform under rotations as components of an irreducible spherical tensor operator of rank l, specifically the q = m component T_m^{(l)}, following the representation Y_l^m(\hat{r}') = \sum_{m'} Y_l^{m'}(\hat{r}) D_{m' m}^{(l)}(\alpha, \beta, \gamma), with D^{(l)} the Wigner D-matrix elements. In position-dependent operators, such as those arising in atomic potentials or electromagnetic interactions, multipole operators are constructed by coupling the radial coordinate with to form irreducible tensors. The general form is r^k Y_q^{(k)}(\hat{r}), which constitutes an irreducible spherical tensor operator of k, where r is the radial distance and \hat{r} the . These operators describe the of charge or in multipole expansions, with k determining the order (e.g., for k=1, for k=2). The Wigner-Eckart theorem facilitates the evaluation of matrix elements of these operators between states with definite orbital , such as hydrogenic orbitals |n l m\rangle. Specifically, the angular momentum matrix element is \langle l' m' | r^k Y_q^{(k)} | l m \rangle = \langle l m \, k q | l' m' \rangle \langle l' || r^k Y^{(k)} || l \rangle, where \langle l m \, k q | l' m' \rangle is the Clebsch-Gordan coefficient enforcing selection rules |l - k| \leq l' \leq l + k and m' = m + q, and the reduced matrix element \langle l' || r^k Y^{(k)} || l \rangle factors into a radial \int R_{n' l'}(r) r^k R_{n l}(r) r^2 dr multiplied by an angular factor depending only on l, l', k. This decomposition separates the angular coupling from the radial dynamics, simplifying calculations in multi-electron atoms. Orbital states provide the basis for these integrals, assuming separation of radial and angular parts in the wave function. A prominent example in is the electric multipole moments, particularly the dipole operator for k=1, given by \mathbf{d} = -e \mathbf{r}, whose spherical components are proportional to r Y_1^m(\hat{r}) (up to normalization factors like \sqrt{4\pi/3}). These describe transitions between atomic states driven by , such as in optical or . The parity of the multipole operator r^k Y_q^{(k)}(\hat{r}) under spatial inversion is (-1)^k, arising from the even parity of r^k and the intrinsic (-1)^k of the spherical harmonic Y_q^{(k)}, which determines whether the operator connects states of even or odd .

Tensor Operators in Spin Systems

In , the spin angular momentum of a particle is represented by the spin operators \mathbf{S} = (S_x, S_y, S_z), which form a vector operator of transforming under rotations according to the vector of the rotation group SO(3). These operators satisfy the commutation relations [S_i, S_j] = i \hbar \epsilon_{ijk} S_k for i, j, k = x, y, z, identical to those of orbital operators, confirming their role as generators of rotations in spin space. For systems involving both orbital and spin degrees of freedom, the total angular momentum operator is defined as the vector sum \mathbf{J} = \mathbf{L} + \mathbf{S}, where \mathbf{L} is the orbital angular momentum operator, also a rank-1 vector operator. While \mathbf{J} itself is a vector operator, higher-rank tensor operators in spin systems arise from tensor products of \mathbf{S} with itself or with other operators, enabling the description of interactions beyond simple vector couplings; are used to decompose these products into irreducible tensor components when coupling spins to form total \mathbf{J}. A key application of tensor operators in spin systems is the spin-orbit interaction, captured by the Hamiltonian term \lambda \mathbf{L} \cdot \mathbf{S}, where \lambda is the spin-orbit coupling constant. This , a scalar (rank-0 tensor) formed from two vector operators, can be rewritten using the \mathbf{L} \cdot \mathbf{S} = \frac{1}{2} (J^2 - L^2 - S^2), yielding shifts that depend on the quantum numbers of \mathbf{J}, \mathbf{L}, and \mathbf{S}: \lambda \mathbf{L} \cdot \mathbf{S} = \frac{\lambda}{2} [J(J+1) - L(L+1) - S(S+1)] \hbar^2. This form highlights how the interaction mixes orbital and spin contributions into effective states. Higher-rank tensor operators are essential for describing environmental effects on states, such as crystal potentials in ionic solids, which split degenerate levels through electrostatic interactions. For ions with partially filled d-shells, the crystal Hamiltonian is expanded in irreducible tensor operators of even ranks up to 4 (ranks 2 and 4), acting on the and orbital states to produce anisotropic splittings. In like iron or compounds, this leads to an effective Hamiltonian incorporating tensor shifts, such as the rank-2 zero-field splitting parameter D and the anisotropic g-tensor, which account for deviations from isotropic behavior due to the local crystal symmetry. These tensor terms enable precise modeling of magnetic properties, with parameters fitted to experimental spectra to reveal -orbit and crystal influences.

Applications

Dipole Transitions in Atoms

In , the interaction between a single-electron atom and an under the electric is governed by the perturbation H' = -\vec{d} \cdot \vec{E}, where \vec{d} = -e \vec{r} is the operator for the valence electron and \vec{E} is the strength. This Hamiltonian arises from expanding the full interaction potential and retaining the leading term for wavelengths much larger than atomic dimensions. The \vec{r} acts as a (rank-1) irreducible tensor operator, enabling the use of spherical tensor to evaluate elements. The probability amplitude for an electric dipole transition from an initial state |i\rangle to a final state |f\rangle is determined by the matrix element \langle f | d_q^{(1)} | i \rangle, where d_q^{(1)} are the spherical components of the dipole operator. According to Fermi's golden rule, the transition rate w_{i \to f} is given by w_{i \to f} = \frac{2\pi}{\hbar} |\langle f | H' | i \rangle|^2 \delta(E_f - E_i - \hbar \omega), which for dipole transitions simplifies to a form proportional to |\langle f | d_q^{(1)} | i \rangle|^2 times the field intensity and density of states. The Wigner-Eckart theorem decomposes this matrix element as \langle \alpha' l' m' | d_q^{(1)} | \alpha l m \rangle = \langle l m 1 q | l' m' \rangle \frac{\langle \alpha' l' || d^{(1)} || \alpha l \rangle}{\sqrt{2l' + 1}}, where \alpha denotes other quantum numbers (e.g., principal quantum number n), the Clebsch-Gordan coefficient \langle l m 1 q | l' m' \rangle encodes the angular dependence, and \langle \alpha' l' || d^{(1)} || \alpha l \rangle is the reduced matrix element independent of magnetic quantum numbers. This separation highlights how tensor operators factorize the transition into angular (from spherical harmonics coupling) and radial contributions. The Clebsch-Gordan coefficients impose selection rules on allowed transitions: \Delta l = \pm 1 for the orbital angular momentum quantum number and \Delta m = 0, \pm 1 depending on the polarization q. Additionally, the odd parity of the dipole operator requires a change in the parity of the wave functions, ensuring transitions occur only between states of opposite parity (e.g., from even to odd l). For single-electron atoms such as alkali metals, where the valence electron dominates, the reduced matrix element reduces to \langle n' l' || d^{(1)} || n l \rangle = -e \langle R_{n' l'} | r | R_{n l} \rangle \times (angular factor from recoupling), with the radial integral \int_0^\infty R_{n' l'}(r) r R_{n l}(r) r^2 dr capturing the overlap of radial wave functions R_{nl}. A representative example is the $2p \to 2s transition in hydrogen-like atoms, where the non-zero radial integral \langle R_{2p} | r | R_{2s} \rangle (approximately \sqrt{24} a_0 / Z for nuclear charge Z) confirms the transition's allowance despite degeneracy in energy levels. The intensity of spectral lines from these transitions is quantified by the line strength S = |\langle n l || d^{(1)} || n' l' \rangle|^2, which is independent of magnetic sublevels and scales with the square of the reduced matrix element, providing a measure of transition probability averaged over polarizations. In alkali atoms, this framework accurately predicts observed optical transitions, such as those in the sodium D lines, by evaluating the radial for the outer while treating the core as a screened potential.

Magnetic Resonance Spectroscopy

In magnetic resonance spectroscopy (MRS), tensor operators provide a framework for describing the anisotropic interactions in spin Hamiltonians, particularly in solid-state NMR and electron paramagnetic resonance (EPR) where molecular orientations lead to direction-dependent effects. These operators, often expressed as irreducible spherical tensors, transform under rotations according to the rules of angular momentum algebra, enabling the analysis of spectra influenced by crystal fields or molecular tumbling. The Zeeman interaction, central to MRS, exemplifies this: in the isotropic lab frame, the Hamiltonian is H_Z = - \boldsymbol{\mu} \cdot \mathbf{B} = g \mu_B S_z B, a scalar product of vector operators, but in crystalline environments, it manifests as a rank-2 tensor due to g-anisotropy. The hyperfine interaction between electron and nuclear spins, H_{hf} = \mathbf{S} \cdot \mathbf{A} \cdot \mathbf{I}, arises from the scalar contraction of rank-1 tensor operators for the electron spin \mathbf{S} and nuclear spin \mathbf{I}, with the hyperfine tensor \mathbf{A} incorporating isotropic Fermi contact, dipolar, and orbital contributions. In EPR, this tensor reveals the local geometry around paramagnetic centers, splitting spectra into multiplets whose patterns depend on the principal values and axes of \mathbf{A}. For nuclei with spin I > 1/2, the quadrupolar Hamiltonian H_Q \propto \sum_{i,j} Q_{ij} I_i I_j represents a rank-2 tensor interaction between the nuclear electric quadrupole moment and the electric field gradient, dominating spectra in solid-state NMR of quadrupolar nuclei like ^{23}\mathrm{Na} or ^{27}\mathrm{Al}. This term shifts energy levels and broadens lines, with the coupling constant C_Q and asymmetry \eta quantifying the tensor's principal components. Relaxation processes in MRS, governing longitudinal (T_1) and transverse (T_2) times, stem from fluctuating tensor operators correlated with lattice motions, analyzed via Redfield theory through spectral densities J(\omega) at Larmor frequencies. Dipolar and quadrupolar mechanisms involve rank-1 and rank-2 tensor correlations, where $1/T_1 \propto J(\omega_I + \omega_S) + J(\omega_I - \omega_S) for heteronuclear cases, linking rates to on picosecond to nanosecond scales. In EPR of solids, the g-tensor, a rank-2 describing Zeeman , shifts resonance fields according to g = g_e (1 + \Delta g), with principal values revealing spin-orbit and in materials like defects. This broadens powder patterns, enabling structural insights into paramagnetic species.

Selection Rules in Quantum Transitions

The Wigner–Eckart theorem provides a fundamental framework for evaluating matrix elements of irreducible tensor operators in quantum mechanics, separating the angular momentum dependence from the intrinsic dynamical factors. Specifically, for a spherical tensor operator T_q^{(k)} acting between angular momentum states |\gamma j m\rangle and |\gamma' j' m'\rangle, where \gamma denotes additional quantum numbers, the theorem states: \langle \gamma' j' m' | T_q^{(k)} | \gamma j m \rangle = \langle j m, k q | j' m' \rangle \langle \gamma' j' || T^{(k)} || \gamma j \rangle, with the Clebsch–Gordan coefficient \langle j m, k q | j' m' \rangle encapsulating the angular part and the reduced matrix element \langle \gamma' j' || T^{(k)} || \gamma j \rangle independent of the magnetic quantum numbers m, m', q. This decomposition arises from the representation theory of the rotation group SU(2), ensuring rotational invariance. The selection rules for quantum transitions induced by such operators follow directly from the orthogonality properties of the Clebsch–Gordan coefficients. The matrix element vanishes unless m' = m + q, enforcing conservation of the z-component of angular momentum up to the operator's projection q. For the total angular momentum quantum numbers, the triangle inequality must hold: |j - j'| \leq k \leq j + j', which implies \Delta j = 0, \pm 1, \dots, \pm k. Additionally, transitions from j = 0 to j' = 0 are forbidden for k \geq 1, as no non-zero Clebsch–Gordan coefficient exists in that case, though k = 0 scalars allow $0 \to 0 connections. These rules generalize the familiar dipole selection rules (where k = 1) to higher-rank multipole operators. Parity provides an additional , as tensor operators possess definite under spatial inversion. For a polar tensor operator of k, the is (-1)^k, requiring the product of the parities of the initial and final states to match that of the operator for the matrix element to be non-zero; thus, \Delta \pi = (-1)^k, where \Delta \pi = \pi' / \pi and \pi, \pi' are the state parities. Axial tensors, such as those for magnetic interactions, have opposite (-1)^{k+1}. This leads to forbidden transitions when is violated; for instance, in systems with definite orbital (-1)^l, electric (k=1, odd ) forbids \Delta l = 0 (even change), while magnetic (k=1, even ) allows \Delta l = 0 or even changes, enabling otherwise forbidden lines. The dynamical strength of allowed transitions is encoded in the reduced matrix element, which determines transition rates and lifetimes. For electric multipole transitions, the oscillator strength f, a measure of transition probability, is proportional to the square of the reduced matrix element averaged over initial degeneracies: f \propto |\langle \gamma' j' || T^{(k)} || \gamma j \rangle|^2 / (2j + 1). This quantity, independent of angular details, captures the radial or intrinsic overlap, allowing lifetime estimates via \tau^{-1} \propto \omega^3 f for radiative decay, where \omega is the transition frequency.

References

  1. [1]
    4.8: Tensor Operators - Physics LibreTexts
    Mar 5, 2022 · A tensor is a generalization of a such a vector to an object with more than one suffix with the requirement that these components mix among ...The Rotation Operator in... · Rotating a Basis Ket · Vector Operators: Definition...
  2. [2]
    [PDF] Vector and Tensor operators in quantum mechanics - bingweb
    Sep 3, 2017 · We discuss the properties of vector and tensor operators under rotations in quantum. mechanics. We use the analogy from the classical physics, ...
  3. [3]
    [PDF] Tensor operators
    Tensor operators. We have used the vector notation for three component operators for example to express the scalar product, like.
  4. [4]
  5. [5]
    Tensor Operators - Galileo
    A tensor is a generalization of a vector with more than one suffix, where components mix under rotation. The number of suffixes is the rank of the tensor.
  6. [6]
    [PDF] Interpreting the Wigner–Eckart theorem - Josh R. Hunt
    In his 1931 monograph on applications of group theory to atomic spectra, Wigner was the first to state the Wigner–Eckart theorem for arbitrary tensor operators ...
  7. [7]
    [PDF] A User's Guide to Angular Momentum Theory - MIT OpenCourseWare
    • Commutation rule definition of a spherical tensor operator, T. µ k (k rank, µ component,. –k ≤ µ ≤ k: scalar k=0, vector k=1, and k=2 2nd rank tensor).
  8. [8]
    [PDF] Physics 221A Fall 1996 Notes 15 Irreducible Tensor Operators and ...
    Tensor operators of other ranks (besides 2) are possible; a scalar is considered a tensor operator of rank 0, and a vector is considered a tensor of rank 1. In ...
  9. [9]
    [PDF] Irreducible Tensor Operators and the Wigner-Eckart Theorem
    Similarly, in quantum mechanics, we define a vector operator as a vector of operators (that is, a set of three operators) with certain transformation properties ...
  10. [10]
    [PDF] Rotations and Angular Momentum - CMU Quantum Theory Group
    The unitary quantum mechanical operator which represents an active rotation ~ω can be written as. R(~ω) = exp[−i~ω · ~J],. (10) where ~ω · ~J = ωxJx + ωyJy + ...
  11. [11]
    [PDF] Rotations in Quantum Mechanics, and Rotations of Spin-1
    These experiments are discussed in more detail by. Sakurai, Modern Quantum Mechanics, and they show that we must take the −1 phase shift for 2π rotations to be ...
  12. [12]
    [PDF] Quantum Mechanics - IFSC/USP
    3.3.1 The orbital angular momentum operator. The definition of orbital angular momentum is adopted from classical mechanics: L = r × p = −ıhr × ∇ = −ıh êx.
  13. [13]
    [PDF] Representations of the Angular Momentum Operators and Rotations
    In Notes 10, the angular momentum J of a quantum system was defined by Eq. (10.7) as the infinitesimal generator of rotations, and it was shown that the ...
  14. [14]
    Vector and tensor operators - Book chapter - IOPscience
    The vector and tensor character of operators is described. This is developed into identifiable commutator bracket relationships with the operator components ...
  15. [15]
  16. [16]
    [PDF] ANGULAR MOMENTUM - ResearchGate
    May 2, 2017 · Angular Momentum in Quantum Mechanics by A. R. Edmonds. Page 4 ... The Tensor Operators in Quantum Mechanics............... 71. 5.3 ...
  17. [17]
    [PDF] Quantum Mechanics - Universidad de Granada
    Oct 15, 2025 · Vector operators have been defined as satisying [Ji,Vj] = ihϵijkVk. Let's introduce spherical coordinates: V + ≡ −. V1 + iV2. √2. ,V ...
  18. [18]
    [PDF] 1 Notes on spherical tensors and Wigner-Eckart theorem
    So each index transforms as a vector under the rotation group. The simplest example is that of a Cartesian rank 2 tensor constructed out of two vectors. Wij = ...
  19. [19]
    [PDF] Magnetism, Angular Momentum, and Spin - Grandinetti Group
    Coupling of e−'s magnetic dipole moment to external magnetic field is called Zeeman interaction. Taking magnetic field as along z axis, ⃗B = Bz ⃗ez, expression ...
  20. [20]
    [PDF] Note on electric multipole expansions in quantum mechanics
    Recall that in classical electrostatics it is often useful to express electric potentials as multipole expansions, i.e. in terms of spherical harmonics ...
  21. [21]
    [PDF] 5.74 rwf lecture #3 rotational transformations and spherical tensor ...
    Spherical tensor expansion is like a multipole expansion. Anything can be broken up into angular momentum-like parts, including what a laser writes onto a ...
  22. [22]
    [PDF] The tensor spherical harmonics
    The basics of angular momentum theory, Clebsch-Gordon coefficients, spherical tensors and the Wigner-Eckart theorem are treated in most quantum mechanics ...
  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]
  25. [25]
  26. [26]
    Angular Momentum Operators - Richard Fitzpatrick
    The operator $L^2$, which represents the magnitude squared of the angular momentum vector, is defined \begin{displaymath} L^2 = L_x^{ 2} + L_y^{ 2} + L_z^{ 2}.
  27. [27]
    [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 ...
  28. [28]
    Angular Momentum Operator Algebra - Galileo
    We have therefore established that the orbital angular momentum operator ˆ→L is the generator of spatial rotations, by which we mean that if we rotate our ...
  29. [29]
    Quantum Numbers and Electron Configurations
    The angular quantum number (l) describes the shape of the orbital. Orbitals have shapes that are best described as spherical (l = 0), polar (l = 1), or ...
  30. [30]
    [PDF] Atomic Physics lectures University of Amsterdam
    Dec 1, 2018 · ... spherical basis . . . . . . . . . . . . . . . . . . . . 29. 2.5.3 ... vector operators . . . . . . . 38. 3.2 Matrix representation of ...
  31. [31]
    [PDF] Part III quantum mechanics preparatory workshop notes - DAMTP
    We now create a new set of three operators, ˆSi, called spin operators and ... for the components of the spin vector (. ˆ. ⃗S)i = ˆSi. So far, this is an ...
  32. [32]
    [PDF] Today's Outline - February 04, 2020
    Feb 4, 2020 · This can be used to recast the spin-orbit Hamiltonian. J2 = (L + S) · (L + S) = L2 + S2 + 2L ·. S. L · S = 1. 2. (J2 − L2 − S2). C. Segre (IIT).
  33. [33]
    Spin-Hamiltonian parameters of ions | Phys. Rev. B
    Jun 1, 1988 · The importance of utilizing the appropriate crystal-field model is stressed, and the observed axial term D of and ions at the tetrahedral and ...
  34. [34]
    [PDF] 5.74 Introductory Quantum Mechanics II - MIT OpenCourseWare
    Here, we'll derive the Hamiltonian for the light-matter interaction, the Electric Dipole. Hamiltonian. It is obtained by starting with the force experienced ...
  35. [35]
    [PDF] Equivalence of interaction Hamiltonians in the electric dipole ...
    Abstract. The interaction of an atomic system with an externally applied electromagnetic field can be treated in the electric dipole approximation.
  36. [36]
    [PDF] Tensor Operators and the Wigner Eckart Theorem
    Nov 11, 2009 · Vector operators transform under rotation. Cartesian tensors are formed by combining two vector operators. A spherical tensor transforms ...
  37. [37]
    [PDF] arXiv:0804.4528v1 [physics.atom-ph] 29 Apr 2008
    Apr 29, 2008 · In Sec. IIIB, I explain how the Wigner-Eckart theorem can simplify calculations for transitions between states of well-defined angular mo-.<|control11|><|separator|>
  38. [38]
    [PDF] Atoms in a Radiation field Masatsugu Sei Suzuki ... - bingweb
    Feb 14, 2017 · absorption using the Fermi's golden rule and the Wigner-Eckart theorem. Both the stimulated emission, and absorption are proportional to the ...
  39. [39]
    [PDF] Wigner–Eckart Theorem
    The Wigner-Eckart theorem concerns matrix elements of operators between states with definite angular momentum, reducing the number of independent matrix ...Missing: physics Y_1^
  40. [40]
    [PDF] 8 Electron orbit in atoms - 8.3 Transition selection rule - Xie Chen
    This is the content of the so called Wigner-Eckart theorem. Tensor Operator: An irreducible tensor operator T j m is a set of operators labelled by fixed ...
  41. [41]
    Atomic Spectros. - Spectral Lines - NIST
    Selection rules for discrete transitions. Electric dipole (E1) ("allowed"), Magnetic dipole (M1) ("forbidden"), Electric quadrupole (E2) ("forbidden") ...
  42. [42]
    [PDF] 7.1 Dipole Transitions
    Dipole transitions are commonly used in atomic physics for laser cooling, state preparation, and state measurement.Missing: tensor | Show results with:tensor
  43. [43]
    Transitions in Hydrogen - Richard Fitzpatrick
    $$2P\rightarrow 1S$ Transitions in Hydrogen. Let us calculate the rate of spontaneous emission between the first excited state (i.e., $n=2$ ) and the ground ...
  44. [44]
    [PDF] Lecture 19, Hydrogen Atom - DSpace@MIT
    We expect to see a 2p→1s transition, but not a 2s→1s transition, and this ... For the radial part, we need the integral π. ∫Rnl ( )r rR. 2 n'l' ( )r r ...
  45. [45]
    Generalized Tables for the Calculation of Dipole Transition ...
    Application of the algebra of tensor operators developed by Racah breaks the matrix-elements into a line factor, a multiplet factor, and a transition integral.
  46. [46]
    Electric dipole matrix elements for the transition in atomic cesium
    Mar 5, 2019 · ... rate on polarization by examining the two-photon transition rate expressed through the Fermi golden rule. S = 2 π ℏ | A 2 P | 2 ρ 7 s ( E ) ,.
  47. [47]
    [PDF] 4 Effective Hamiltonians in Chemistry - cond-mat.de
    3.2 The spin Hamiltonian in EPR and NMR spectroscopy. The spin Hamiltonian (SH) is a major asset in the analysis of magnetic resonance experiments. The ...
  48. [48]
    Evolution of Tesseral and Spherical Tensor Operators in EPR and ...
    Jun 29, 2025 · Importantly, the ESO [19,42] with the rank k = 2, 4, 6 are sufficient for single transition ions, which possess the highest spin S = 7/2, like ...
  49. [49]
    A master equation for spin systems far from equilibrium - ScienceDirect
    In high-field NMR, the dominant Hamiltonian H A is usually identified with the Zeeman Hamiltonian ... spherical tensor operator sets:(51) { A α } = { T λ ...
  50. [50]
    Quadratic Spin–Orbit Mechanism of the Electronic g-Tensor
    Mar 10, 2023 · An implementation of the g-tensor of EPR spectroscopy is presented. This implementation is based on d. functional theory (DFT) and the use of ...
  51. [51]
    Accurate hyperfine tensors for solid state quantum applications
    Jun 4, 2024 · The two dominant contributions to the hyperfine tensor are the Fermi contact interaction, {{{{{{{{\mathcal{A}}}}}}}}}_{{{{{{{{\rm{FC}}}}}}}}} , ...
  52. [52]
    Determination of nuclear quadrupolar parameters using singularities ...
    Oct 6, 2016 · H Q is the quadrupolar Hamiltonian, which is expressed as a scalar product of second-rank tensors A and T2 as22,23. H Q = h C Q 2 I ( 2 ...
  53. [53]
    23Na relaxometry: An overview of theory and applications
    ... relaxation pathway and the spectral densities that are affected by the dynamics. ... relaxation pathways because they possess different tensor operators ...
  54. [54]
    [PDF] Spectral densities and nuclear spin relaxation in solids - CORE
    "Spectral densities and nuclear spin relaxation in solids." Physics Reports 171. 3: 85-128. Page 2. SPECTRAL DENSITIES AND NUCLEAR SPIN. RELAXATION IN SOLIDS.
  55. [55]
    Electronic relaxation of paramagnetic metal ions and NMR relaxivity ...
    Introducing the tensor operators S q 1 ( ... The numerical values of the normalized electronic TCFs were used for computing the normalized spectral densities ...
  56. [56]
    First-principles calculation of the EPR tensor in extended periodic ...
    Mar 20, 2006 · The g tensor from EPR spectroscopy is a second-order property and can therefore be evaluated using double perturbation theory. The perturbation ...
  57. [57]
    [PDF] The Wigner-Eckart Theorem - University of California, Berkeley
    It turns out that operators of a certain type, the irreducible tensor operators, are associated with angular momentum quantum numbers and have transformation ...
  58. [58]
    [PDF] IX. The Wigner-Eckart theorem - MIT
    For example, if the operators form a spherical tensor with respect to spin rotations, there need be no special relationship of these operators between orbital.
  59. [59]
    [PDF] 7.8 Tensor Operators in Quantum Mechanics
    The special success of quantum me- chanics can be seen in the prediction of observables of systems consisting of more than two particles, i.e., N = 3, 4, ททท. ...
  60. [60]
    [PDF] Einstein coefficients, cross sections, f values, dipole moments ... - arXiv
    The Einstein A and B coefficients, f values (also called. “oscillator strengths”), and transition dipole moments are all atomic and molecular parameters related ...
  61. [61]
    [PDF] 5.74 RWF Lecture #4 4 – 1 The Wigner-Eckart Theorem
    The Wigner-Eckart. Theorem provides an automatic way of evaluating the angular parts of matrix elements of many important types of operators. The radial factor ...