Fact-checked by Grok 2 weeks ago

Pauli matrices

The Pauli matrices are a set of three 2×2 complex Hermitian matrices that serve as the basis for representing the spin angular momentum operators of particles in . Named after the physicist , who introduced them in the context of , these matrices are denoted as \sigma_x, \sigma_y, and \sigma_z (or collectively as \vec{\sigma}). They are explicitly given by: \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 exhibit several key algebraic properties that make them indispensable in quantum physics. Each Pauli matrix squares to the identity matrix: \sigma_x^2 = \sigma_y^2 = \sigma_z^2 = I. They satisfy the commutation relations [\sigma_x, \sigma_y] = 2i \sigma_z and cyclic permutations, which mirror the algebra of . Additionally, the Pauli matrices are traceless, unitary, and form a complete basis for the of Hermitian matrices, allowing any such to be expressed as a : M = \alpha I + \vec{v} \cdot \vec{\sigma}, where \vec{v} is a real vector. In applications, the Pauli matrices underpin the description of two-state , such as the of or other fermions. The operators are defined as \hat{S}_i = \frac{[\hbar](/page/H-bar)}{2} \sigma_i for i = [x, y](/page/X&Y), [z](/page/Z), with eigenvalues \pm \frac{[\hbar](/page/H-bar)}{2} corresponding to spin-up and spin-down states. They appear in the for the of an in a , H = -\vec{\mu} \cdot \vec{B} = \mu_B \vec{B} \cdot \vec{\sigma}, enabling the analysis of phenomena like and the Stern-Gerlach experiment. Beyond , the Pauli matrices extend to quantum information theory, where they represent operations and Pauli gates in . Their fundamental role highlights the non-commutative nature of quantum observables and the vector-like behavior of in three dimensions.

Definition

Explicit Forms

The Pauli matrices are a set of three 2×2 complex matrices that are Hermitian and unitary, conventionally denoted as \sigma_x, \sigma_y, \sigma_z (or equivalently \sigma_1, \sigma_2, \sigma_3). These matrices were introduced by in 1927 to formulate the of the magnetic , particularly for describing . Their explicit forms in the standard basis are given by \begin{align*} \sigma_x &= \begin{pmatrix} 0 & 1 \ 1 & 0 \end{pmatrix}, \ \sigma_y &= \begin{pmatrix} 0 & -i \ i & 0 \end{pmatrix}, \ \sigma_z &= \begin{pmatrix} 1 & 0 \ 0 & -1 \end{pmatrix}. \end{align*} A common notation treats the Pauli matrices as components of a vector \boldsymbol{\sigma} = (\sigma_x, \sigma_y, \sigma_z), allowing linear combinations such as \boldsymbol{\sigma} \cdot \vec{a} = a_x \sigma_x + a_y \sigma_y + a_z \sigma_z for a real vector \vec{a}.

Basic Properties

The Pauli matrices \sigma_x, \sigma_y, and \sigma_z are each Hermitian, satisfying \sigma_i^\dagger = \sigma_i for i = 1, 2, 3. This property ensures they can represent physical observables in , such as components. Additionally, each Pauli matrix is unitary, as \sigma_i^2 = I, where I is the $2 \times 2 , combined with their Hermiticity implying \sigma_i^\dagger \sigma_i = I. This unitarity follows directly from the squaring relation and allows the matrices to generate rotations in space. The trace of each Pauli matrix vanishes: \operatorname{Tr}(\sigma_i) = 0. This tracelessness is a defining feature that distinguishes them from the in expansions of general operators. The squaring property \sigma_i^2 = I holds individually for each i = 1, 2, 3. This identity underscores their role in simple algebraic structures for two-level . Together, the three Pauli matrices form an orthonormal basis (up to normalization) for the three-dimensional real vector space of $2 \times 2 Hermitian traceless matrices. Any such matrix can be uniquely expressed as a real linear combination of \sigma_x, \sigma_y, and \sigma_z.

Algebraic Properties

Commutation Relations

The Pauli matrices \sigma_j (j = x, y, z) obey the commutation relations [\sigma_j, \sigma_k] = 2i \sum_l \epsilon_{jkl} \sigma_l, where \epsilon_{jkl} denotes the Levi-Civita symbol and the sum runs over l = x, y, z. These relations capture the non-commutativity inherent in the algebra of the matrices and were originally derived in the context of electron spin operators. Explicit instances of the general formula include [\sigma_x, \sigma_y] = 2i \sigma_z, [\sigma_y, \sigma_z] = 2i \sigma_x, and [\sigma_z, \sigma_x] = 2i \sigma_y. These follow from the structure of the , which is totally antisymmetric with \epsilon_{xyz} = 1 and vanishes otherwise. A direct verification for one pair proceeds via . With \sigma_x = \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}, \quad \sigma_y = \begin{pmatrix} 0 & -i \\ i & 0 \end{pmatrix}, compute \sigma_x \sigma_y = \begin{pmatrix} i & 0 \\ 0 & -i \end{pmatrix} = i \sigma_z and \sigma_y \sigma_x = \begin{pmatrix} -i & 0 \\ 0 & i \end{pmatrix} = -i \sigma_z, yielding [\sigma_x, \sigma_y] = 2i \sigma_z. The remaining relations hold by of indices. The commutation relations establish that the Pauli matrices, up to multiplication by the factor i, generate the Lie algebra \mathfrak{su}(2). In particular, the basis elements T_j = -\frac{i}{2} \sigma_j satisfy the standard structure equations [T_j, T_k] = \epsilon_{jkl} T_l, providing the fundamental representation of \mathfrak{su}(2).

Anticommutation Relations

The anticommutation relations for the Pauli matrices \sigma_x, \sigma_y, \sigma_z (often denoted collectively as \sigma_j for j = x, y, z) are expressed as \{ \sigma_j, \sigma_k \} = 2 \delta_{jk} I, where \{A, B\} = AB + BA denotes the anticommutator, \delta_{jk} is the Kronecker delta (equal to 1 if j = k and 0 otherwise), and I is the $2 \times 2 identity matrix. These relations were introduced by Pauli in his formulation of the quantum mechanics of the magnetic electron, where the matrices describe the two-valued degree of freedom associated with electron spin. When j = k, the relation simplifies to \sigma_j^2 = I, reflecting that each Pauli matrix is involutory (its own inverse up to a sign). For example, \{ \sigma_x, \sigma_x \} = 2 \sigma_x^2 = 2I. In contrast, for j \neq k, the anticommutator vanishes, yielding \{ \sigma_x, \sigma_y \} = \sigma_x \sigma_y + \sigma_y \sigma_x = 0, and similarly for the other distinct pairs. These properties arise from the matrices' Hermiticity (\sigma_j^\dagger = \sigma_j) and can be verified through direct computation using their explicit forms: \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}. For instance, multiplying \sigma_x \sigma_y = i \sigma_z and \sigma_y \sigma_x = -i \sigma_z confirms their sum is the . The anticommutation relations play a key role in establishing the Pauli matrices as an for the three-dimensional of $2 \times 2 traceless Hermitian matrices, with respect to the Frobenius (Hilbert-Schmidt) inner product \langle A, B \rangle = \operatorname{Tr}(A^\dagger B). Since the matrices are Hermitian, this inner product reduces to \langle \sigma_j, \sigma_k \rangle = \operatorname{Tr}(\sigma_j \sigma_k) = \frac{1}{2} \operatorname{Tr}(\{ \sigma_j, \sigma_k \}) = \delta_{jk} \operatorname{Tr}(I) = 2 \delta_{jk}, confirming orthogonality for j \neq k and equal norm for each \sigma_j. This orthogonality underpins the unique decomposition of any traceless Hermitian matrix as a real \mathbf{n} \cdot \boldsymbol{\sigma} = n_x \sigma_x + n_y \sigma_y + n_z \sigma_z, where \mathbf{n} is a real vector, facilitating representations in such as the for systems.

Eigenvalues and Eigenvectors

The Pauli matrices \sigma_x, \sigma_y, and \sigma_z each possess eigenvalues +1 and -1. These eigenvalues arise from the \det(\sigma_i - \lambda I) = -\lambda^2 + 1 = 0, which simplifies to \lambda^2 = 1, yielding the solutions \lambda = \pm 1 for each i = x, y, z. The corresponding eigenvectors for each Pauli matrix form an orthonormal basis for \mathbb{C}^2. For \sigma_z, the eigenvector for eigenvalue +1 is |+\rangle = \begin{pmatrix} 1 \\ 0 \end{pmatrix} and for -1 is |-\rangle = \begin{pmatrix} 0 \\ 1 \end{pmatrix}. Similarly, for \sigma_x, the eigenvectors are \frac{1}{\sqrt{2}} \begin{pmatrix} 1 \\ 1 \end{pmatrix} (eigenvalue +1) and \frac{1}{\sqrt{2}} \begin{pmatrix} 1 \\ -1 \end{pmatrix} (eigenvalue -1); for \sigma_y, they are \frac{1}{\sqrt{2}} \begin{pmatrix} 1 \\ i \end{pmatrix} (eigenvalue +1) and \frac{1}{\sqrt{2}} \begin{pmatrix} 1 \\ -i \end{pmatrix} (eigenvalue -1). These eigenvectors are defined up to an overall phase factor. More generally, the operator \vec{\sigma} \cdot \hat{n}, where \hat{n} is a , also has eigenvalues \pm [1](/page/1). Its eigenvectors correspond to the spin-up and spin-down states aligned along the \pm \hat{n} directions, respectively. In the \sigma_z basis, these take the form |+\hat{n}\rangle = \begin{pmatrix} e^{-i\phi/2} \cos(\theta/2) \\ e^{i\phi/2} \sin(\theta/2) \end{pmatrix} for +[1](/page/1) and |-\hat{n}\rangle = \begin{pmatrix} -e^{-i\phi/2} \sin(\theta/2) \\ e^{i\phi/2} \cos(\theta/2) \end{pmatrix} for -[1](/page/1), with \theta and \phi the polar and azimuthal angles of \hat{n}. The spectral decomposition of a Pauli matrix \sigma_i yields the projection operators P_+ = \frac{I + \sigma_i}{2} (onto the +1 eigenspace) and P_- = \frac{I - \sigma_i}{2} (onto the -1 eigenspace). These projectors satisfy P_+ + P_- = I and P_+ P_- = 0, reflecting the completeness of the eigenbasis.

Pauli Vector Formalism

Definition and Operations

In the Pauli vector formalism, the Pauli matrices \sigma_x, \sigma_y, \sigma_z are collected into a \vec{\sigma} = (\sigma_x, \sigma_y, \sigma_z). This notation facilitates the representation of linear combinations as \vec{\sigma} \cdot \vec{a} = a_x \sigma_x + a_y \sigma_y + a_z \sigma_z, where \vec{a} = (a_x, a_y, a_z) is a real three-dimensional , often taken as a for simplicity in applications. The explicit forms of the Pauli matrices are \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 for i = x, y, z, ensuring that \vec{\sigma} \cdot \vec{a} is Hermitian when \vec{a} is real. A key property of the Pauli vector combination is its squaring relation: (\vec{\sigma} \cdot \vec{a})^2 = |\vec{a}|^2 I, where I is the $2 \times 2 identity matrix. For a unit vector \vec{a} with |\vec{a}| = 1, this simplifies to (\vec{\sigma} \cdot \vec{a})^2 = I. This involutory behavior mirrors aspects of unit vectors in three-dimensional Euclidean space and underscores the analogy between the Pauli vector and classical vector algebra. The operator \vec{\sigma} \cdot \vec{a} has eigenvalues \pm 1 and corresponding eigenvectors that represent spin-up and spin-down states along the direction \vec{a}, as detailed in the section on eigenvalues and eigenvectors. The Pauli vector formalism exhibits vector-like multiplication rules, particularly through the product identity (\vec{\sigma} \cdot \vec{a})(\vec{\sigma} \cdot \vec{b}) = \vec{a} \cdot \vec{b} \, I + i (\vec{a} \times \vec{b}) \cdot \vec{\sigma}, where \vec{a} \cdot \vec{b} is the scalar dot product and \vec{a} \times \vec{b} is the vector cross product. This decomposes the matrix product into a scalar multiple of the identity plus an imaginary vector term, capturing both parallel and perpendicular components of \vec{a} and \vec{b}. From this, the commutation relation follows as [\vec{\sigma} \cdot \vec{a}, \vec{\sigma} \cdot \vec{b}] = 2i (\vec{a} \times \vec{b}) \cdot \vec{\sigma}, which equals twice the imaginary part of the product and aligns with the Lie algebra structure of rotations in quantum mechanics. These identities enable efficient computations in spin systems by translating three-dimensional vector operations into $2 \times 2 matrix algebra.

Trace and Determinant Identities

The Pauli matrices satisfy several fundamental trace identities that arise from their algebraic structure, particularly the anticommutation relations \{\sigma_i, \sigma_j\} = 2 \delta_{ij} I and commutation relations [\sigma_i, \sigma_j] = 2i \epsilon_{ijk} \sigma_k, where I is the $2 \times 2 , \delta_{ij} is the , and \epsilon_{ijk} is the . Taking the trace of the anticommutation relation yields \operatorname{Tr}(\sigma_i \sigma_j + \sigma_j \sigma_i) = 2 \delta_{ij} \operatorname{Tr}(I) = 4 \delta_{ij}. Since the trace is cyclic, \operatorname{Tr}(\sigma_i \sigma_j) = \operatorname{Tr}(\sigma_j \sigma_i), it follows that $2 \operatorname{Tr}(\sigma_i \sigma_j) = 4 \delta_{ij}, so \operatorname{Tr}(\sigma_i \sigma_j) = 2 \delta_{ij}. This identity highlights the orthonormality of the Pauli matrices under the Hilbert-Schmidt inner product \langle A, B \rangle = \operatorname{Tr}(A^\dagger B), as they are Hermitian. For the product of three Pauli matrices, the identity \operatorname{Tr}(\sigma_i \sigma_j \sigma_k) = 2i \epsilon_{ijk} can be derived using the \sigma_i \sigma_j = \delta_{ij} I + i \epsilon_{ijl} \sigma_l. Multiplying by \sigma_k gives \sigma_i \sigma_j \sigma_k = \delta_{ij} \sigma_k + i \epsilon_{ijl} \sigma_l \sigma_k. Taking the trace, the first term vanishes because \operatorname{Tr}(\sigma_k) = 0, while the second term simplifies using \sigma_l \sigma_k = \delta_{lk} I + i \epsilon_{lk m} \sigma_m, leading to \operatorname{Tr}(\sigma_i \sigma_j \sigma_k) = i \epsilon_{ijl} \delta_{lk} \operatorname{Tr}(I) = 2i \epsilon_{ijk} after relabeling indices and using \operatorname{Tr}(I) = 2. This relation encodes the totally antisymmetric structure of the Pauli and is useful for computing higher-order invariants. In the Pauli vector formalism, the combination (\boldsymbol{\sigma} \cdot \mathbf{a})(\boldsymbol{\sigma} \cdot \mathbf{b}) = \mathbf{a} \cdot \mathbf{b}\, I + i (\mathbf{a} \times \mathbf{b}) \cdot \boldsymbol{\sigma} leads directly to the identity \operatorname{Tr}[(\boldsymbol{\sigma} \cdot \mathbf{a})(\boldsymbol{\sigma} \cdot \mathbf{b})] = 2 \mathbf{a} \cdot \mathbf{b}, as the cross-product term is traceless. This follows from explicit expansion or from the two-index , since \operatorname{Tr}[(\boldsymbol{\sigma} \cdot \mathbf{a})(\boldsymbol{\sigma} \cdot \mathbf{b})] = \sum_{i,j} a_i b_j \operatorname{Tr}(\sigma_i \sigma_j) = 2 \sum_i a_i b_i = 2 \mathbf{a} \cdot \mathbf{b}. The determinant of a Pauli vector combination satisfies \det(\boldsymbol{\sigma} \cdot \mathbf{a}) = -|\mathbf{a}|^2. This can be shown by noting that (\boldsymbol{\sigma} \cdot \mathbf{a})^2 = |\mathbf{a}|^2 I, so the eigenvalues \lambda obey \lambda^2 = |\mathbf{a}|^2, yielding \lambda = \pm |\mathbf{a}| (with the sign choice ensuring the product is negative due to the odd dimension of the representation). Thus, the determinant, as the product of eigenvalues, is -|\mathbf{a}|^2. For a unit vector \mathbf{a} with |\mathbf{a}| = 1, this simplifies to \det(\boldsymbol{\sigma} \cdot \mathbf{a}) = -1. These identities can also be verified through explicit computation using the standard representations of the Pauli matrices, confirming their consistency across algebraic and matrix-based approaches.

Completeness Relation

The Pauli matrices, together with the 2×2 , form a complete basis for the of 2×2 complex matrices. Specifically, denoting \sigma_0 = I and \sigma_i for i = 1, 2, 3 as the standard Pauli matrices, any 2×2 matrix A can be uniquely expanded as A = \sum_{k=0}^3 a_k \sigma_k, where the coefficients are given by a_k = \frac{1}{2} \operatorname{Tr}(A \sigma_k). This expansion is particularly useful for Hermitian matrices, which span a 4-dimensional real . For any 2×2 Hermitian matrix A, the coefficients a_k are real, and the decomposition simplifies to A = \frac{1}{2} \operatorname{Tr}(A) \, I + \frac{1}{2} \sum_{i=1}^3 \operatorname{Tr}(A \sigma_i) \, \sigma_i. If A is traceless (i.e., \operatorname{Tr}(A) = 0), then A = \frac{1}{2} \sum_{i=1}^3 \operatorname{Tr}(A \sigma_i) \, \sigma_i. This property underscores the role of the Pauli matrices as a basis for traceless Hermitian operators in , such as spin observables. The set \{\sigma_0, \sigma_1, \sigma_2, \sigma_3\} is orthonormal with respect to the Hilbert-Schmidt inner product defined by \langle B, C \rangle = \frac{1}{2} \operatorname{Tr}(B^\dagger C), satisfying \langle \sigma_j, \sigma_k \rangle = \delta_{jk}. This orthonormality follows from the trace identities \operatorname{Tr}(\sigma_j \sigma_k) = 2 \delta_{jk}, which hold because the Pauli matrices are Hermitian (\sigma_k^\dagger = \sigma_k) and their products yield twice the identity for matching indices while off-diagonals average to zero. The completeness of this basis can be proven using trace orthogonality. Assume A = \sum_{k=0}^3 c_k \sigma_k; taking the inner product with \sigma_j gives \langle A, \sigma_j \rangle = \sum_k c_k \langle \sigma_k, \sigma_j \rangle = c_j, so c_j = \frac{1}{2} \operatorname{Tr}(A \sigma_j). Since the four matrices are linearly independent (as evidenced by their in a 4-dimensional space) and span the full space of matrices, the expansion is unique and complete.

Group-Theoretic Aspects

Exponential Forms

The exponential map for elements of the SU(2) generated by the Pauli matrices takes a particularly simple closed form when exponentiating a Pauli vector. For a \mathbf{n} \in \mathbb{R}^3, the matrix exponential is given by \exp\left(i \frac{\theta}{2} \boldsymbol{\sigma} \cdot \mathbf{n}\right) = \cos\left(\frac{\theta}{2}\right) I + i \sin\left(\frac{\theta}{2}\right) (\boldsymbol{\sigma} \cdot \mathbf{n}), where \theta \in \mathbb{R} is the angle parameter, I is the $2 \times 2 , and \boldsymbol{\sigma} = (\sigma_x, \sigma_y, \sigma_z) denotes the vector of Pauli matrices. This formula parametrizes all elements of SU(2) via the choice of \mathbf{n} and \theta. The derivation relies on the power series expansion of the matrix exponential and the key algebraic property (\boldsymbol{\sigma} \cdot \mathbf{n})^2 = I for |\mathbf{n}| = 1. Substituting into the series \exp(X) = \sum_{k=0}^\infty \frac{X^k}{k!} with X = i \frac{\theta}{2} (\boldsymbol{\sigma} \cdot \mathbf{n}) yields even powers that reduce to \cos(\theta/2) I and odd powers that reduce to i \sin(\theta/2) (\boldsymbol{\sigma} \cdot \mathbf{n}), as higher powers cycle between I and \boldsymbol{\sigma} \cdot \mathbf{n}. A significant group-theoretic implication arises from the adjoint action of these exponentials on another Pauli vector \boldsymbol{\sigma} \cdot \mathbf{m}, where \mathbf{m} \in \mathbb{R}^3. The conjugation formula is \exp\left(i \frac{\theta}{2} \boldsymbol{\sigma} \cdot \mathbf{n}\right) (\boldsymbol{\sigma} \cdot \mathbf{m}) \exp\left(-i \frac{\theta}{2} \boldsymbol{\sigma} \cdot \mathbf{n}\right) = \boldsymbol{\sigma} \cdot \left[ \mathbf{m} \cos \theta + (\mathbf{n} \times \mathbf{m}) \sin \theta + \mathbf{n} (\mathbf{n} \cdot \mathbf{m}) (1 - \cos \theta) \right]. This expression rotates the vector \mathbf{m} by angle \theta around the axis \mathbf{n} in the of SU(2), which is isomorphic to SO(3). The composition law for two such SU(2) elements further highlights the group's structure. Consider U = \cos(\alpha/2) I + i \sin(\alpha/2) (\boldsymbol{\sigma} \cdot \mathbf{u}) and V = \cos(\beta/2) I + i \sin(\beta/2) (\boldsymbol{\sigma} \cdot \mathbf{v}), where \mathbf{u}, \mathbf{v} are unit vectors and \alpha, \beta \in \mathbb{R}. Their product is UV = \cos(\gamma/2) I + i \sin(\gamma/2) (\boldsymbol{\sigma} \cdot \mathbf{w}), with \cos\left(\frac{\gamma}{2}\right) = \cos\left(\frac{\alpha}{2}\right) \cos\left(\frac{\beta}{2}\right) - \sin\left(\frac{\alpha}{2}\right) \sin\left(\frac{\beta}{2}\right) (\mathbf{u} \cdot \mathbf{v}) and \sin\left(\frac{\gamma}{2}\right) \mathbf{w} = \sin\left(\frac{\alpha}{2}\right) \cos\left(\frac{\beta}{2}\right) \mathbf{u} + \cos\left(\frac{\alpha}{2}\right) \sin\left(\frac{\beta}{2}\right) \mathbf{v} + \sin\left(\frac{\alpha}{2}\right) \sin\left(\frac{\beta}{2}\right) (\mathbf{u} \times \mathbf{v}). This closed-form multiplication rule, analogous to composition, ensures the product remains an SU(2) element. These forms underscore SU(2)'s role as a double cover of the rotation group SO(3), with details on the explicit mapping provided in the section on connections to SO(3) and s.

SU(2) Representation

The special unitary group SU(2) consists of all 2×2 complex unitary matrices with determinant equal to one. Its elements can be parametrized using the Pauli matrices as U = \cos(\phi/2) \, I - i \sin(\phi/2) \, (\mathbf{\sigma} \cdot \mathbf{u}), where \phi is a real angle, I is the 2×2 identity matrix, \mathbf{u} is a unit vector in \mathbb{R}^3, and \mathbf{\sigma} = (\sigma_x, \sigma_y, \sigma_z) denotes the vector of Pauli matrices. This form arises from the exponential map of the Lie algebra, as detailed in the section on exponential forms. The Lie algebra \mathfrak{su}(2) of SU(2) is the three-dimensional real vector space of 2×2 traceless anti-Hermitian matrices, with basis elements given by -i \sigma_k / 2 for k = 1, 2, 3. The of \mathfrak{su}(2) are determined by the commutators of these basis elements: [-i \sigma_j / 2, -i \sigma_k / 2] = \epsilon_{jkl} (-i \sigma_l / 2), where \epsilon_{jkl} is the , reflecting the \mathfrak{su}(2) \cong \mathfrak{so}(3). The Pauli matrices provide the fundamental of SU(2), which is two-dimensional and acts on \mathbb{C}^2. In this representation, the generators are \sigma_k / 2, satisfying the algebra [\sigma_j / 2, \sigma_k / 2] = i \epsilon_{jkl} (\sigma_l / 2). This 2-dimensional representation is faithful and defines the sector in . SU(2) is a simply connected that acts as a double cover of the group SO(3), meaning there is a 2-to-1 from SU(2) onto SO(3) such that each in SO(3) corresponds to two elements in SU(2), differing by a sign. This covering property is evident in the parametrization, where a full 360° in SO(3) maps to U = -I in SU(2), requiring a 720° to return to the .

Connections to SO(3) and Quaternions

The Pauli matrices provide a key link between the special unitary group SU(2) and the SO(3) through a surjective \phi: \mathrm{SU}(2) \to \mathrm{SO}(3) with kernel \{I, -I\}, establishing SU(2) as a double cover of SO(3). This homomorphism arises from the adjoint action on the space of Pauli matrices, where for U \in \mathrm{SU}(2), the map \phi(U) acts on a vector \mathbf{v} \in \mathbb{R}^3 by conjugating the corresponding \sum v_i \sigma_i: \phi(U) \mathbf{v} = \frac{1}{2} \mathrm{Tr} \left( \sigma_j U \left( \sum v_i \sigma_i \right) U^\dagger \right). This action preserves the inner product on \mathbb{R}^3, inducing rotations in SO(3). Specifically, for a by \theta around unit \mathbf{n}, the corresponding U = \exp(-i \theta \mathbf{n} \cdot \boldsymbol{\sigma}/2) yields the : R(\theta \mathbf{n}) \mathbf{v} = \cos\theta \, \mathbf{v} + (1 - \cos\theta) (\mathbf{n} \cdot \mathbf{v}) \mathbf{n} + \sin\theta \, (\mathbf{n} \times \mathbf{v}). The connection manifests in the representation of rotations: elements of SU(2) generated by the Pauli matrices correspond to rotations in SO(3) via this adjoint map, but the double cover implies that U and -U produce the same , reflecting the non-trivial topology of SU(2) \simeq S^3. For systems, this leads to the physical consequence that a 360° ($2\pi) of the state transforms it by a of -1, requiring a full 720° ($4\pi) to return to the original state, distinguishing half-integer spin from integer spin representations. The Pauli matrices also relate directly to s, where the imaginary units i, j, k satisfy i^2 = j^2 = k^2 = ijk = -1 and can be represented as -i \sigma_x, -i \sigma_y, -i \sigma_z in matrix form, embedding the quaternion algebra into 2×2 complex matrices. Unit quaternions, with norm 1, form the group SU(2), as they correspond to matrices U = x I - i (y \sigma_x + z \sigma_y + w \sigma_z) where |x|^2 + |y|^2 + |z|^2 + |w|^2 = 1, and multiplication preserves unitarity. This isomorphism highlights how SU(2) parameterizes rotations, with the conjugation action mirroring quaternion-based vector rotations in \mathbb{R}^3. An extension to relativistic contexts defines the Pauli 4-vector \sigma_\mu = (I, \sigma_1, \sigma_2, \sigma_3) (with \mu = 0,1,2,3), where I is the 2×2 identity, providing a basis for Hermitian 2×2 matrices that represent 4-vectors in Minkowski space via x = x^\mu \sigma_\mu. Define also \bar{\sigma}_\mu = (I, -\sigma_1, -\sigma_2, -\sigma_3). Lorentz transformations act as x' = T x T^\dagger for T \in \mathrm{SL}(2,\mathbb{C}), preserving the Minkowski metric g_{\mu\nu} = \frac{1}{2} \mathrm{Tr}(\sigma_\mu \bar{\sigma}_\nu) with signature (+,-,-,-), thus analogizing spatial rotations to boosts and rotations in spacetime.

Applications in Physics

Quantum Mechanics

In non-relativistic quantum mechanics, the Pauli matrices provide the for representing the angular momentum of particles, such as the . The spin operators are defined as S_i = \frac{\hbar}{2} \sigma_i for i = x, y, z, where \hbar is the reduced Planck's constant and \sigma_i are the Pauli matrices. This formulation allows the two-dimensional of a system to capture the intrinsic , with eigenvalues \pm \frac{\hbar}{2} along any quantization axis. introduced these operators in his 1927 paper to incorporate the of the into the , resolving inconsistencies in atomic spectra and Zeeman effects. The commutation relations of the Pauli matrices, [\sigma_x, \sigma_y] = 2i \sigma_z and cyclic permutations, translate directly to the spin operators via [S_x, S_y] = i \hbar S_z (and cyclic), mirroring the general angular momentum algebra. These relations underpin the Heisenberg uncertainty principle for spin components, implying \Delta S_x \Delta S_y \geq \frac{\hbar}{2} |\langle S_z \rangle|, which prohibits simultaneous precise measurements of non-commuting spin directions. Pauli derived these properties to ensure consistency with experimental observations of spin precession in magnetic fields. The algebra also formalizes the non-classical nature of spin, distinguishing it from orbital angular momentum. Spin measurements along a \mathbf{n} are described by the operator \mathbf{S} \cdot \mathbf{n} = \frac{\hbar}{2} \boldsymbol{\sigma} \cdot \mathbf{n}, with eigenstates corresponding to "up" (+\frac{\hbar}{2}) and "down" (-\frac{\hbar}{2}) along \mathbf{n}. For a general state |\psi\rangle, the probability of measuring spin up along \mathbf{n} is P_+ = \frac{1 + \langle \psi | \boldsymbol{\sigma} \cdot \mathbf{n} | \psi \rangle}{2}, reflecting the expectation value of the projection. This probabilistic outcome aligns with the Stern-Gerlach experiment of 1922, where silver atoms split into two beams in an inhomogeneous , demonstrating quantized spin deflection—later formalized using Pauli matrices. The addition of angular momenta for two spin-1/2 particles yields total spin states of S = 1 (triplet, symmetric) or S = 0 (, antisymmetric), governed by Clebsch-Gordan coefficients that decompose the four-dimensional space into irreducible representations. For instance, the total spin-1 states include symmetric combinations like \frac{1}{\sqrt{2}} (|\uparrow\downarrow\rangle + |\downarrow\uparrow\rangle) for m = 0, while the is \frac{1}{\sqrt{2}} (|\uparrow\downarrow\rangle - |\downarrow\uparrow\rangle). These decompositions, essential for multi-particle systems like the , were systematically tabulated using in early treatments.

Relativistic Quantum Mechanics

In relativistic quantum mechanics, the Pauli matrices are integral to the formulation of the Dirac equation, which describes the behavior of spin-1/2 fermions like electrons in a relativistic setting. The Dirac Hamiltonian takes the form H = c \vec{\alpha} \cdot \vec{p} + \beta m c^2, where the 4×4 Dirac matrices \alpha_i and \beta are constructed using 2×2 blocks involving the Pauli matrices \sigma_i (for i = 1, 2, 3) and the 2×2 identity I_2. In the standard Dirac representation, \beta is block-diagonal as \beta = \begin{pmatrix} I_2 & 0 \\ 0 & -I_2 \end{pmatrix}, while the \alpha_i are block-off-diagonal: \alpha_i = \begin{pmatrix} 0 & \sigma_i \\ \sigma_i & 0 \end{pmatrix}. This block structure ensures the Clifford algebra relations \{\alpha_i, \alpha_j\} = 2\delta_{ij} I_4, \{\alpha_i, \beta\} = 0, and \beta^2 = I_4, allowing the equation to incorporate both relativistic kinematics and spin degrees of freedom consistently. The Foldy-Wouthuysen transformation provides a systematic way to block-diagonalize the Dirac , decoupling the positive- and negative-energy sectors and revealing an effective description for low-velocity particles within the relativistic framework. This unitary transformation, expanded perturbatively in powers of $1/m (where m is the fermion mass), transforms the original into one where the upper two components correspond to positive-energy states and the lower to negative-energy states, with off-diagonal terms eliminated order by order. The resulting effective for positive-energy states includes relativistic corrections such as Darwin terms and spin-orbit interactions, where the spin is represented by the Pauli matrices acting on the two-component spinors. This approach not only bridges the to non-relativistic but also elucidates relativistic effects like in the full relativistic context. For the relativistic , the with a potential yields exact solutions that explain the splitting of spectral lines, arising from relativistic corrections including spin-orbit coupling. The spin-orbit interaction term in the effective is V_{SO} = \frac{1}{2m^2 c^2} \frac{1}{r} \frac{dV}{dr} \vec{S} \cdot \vec{L}, where \vec{S} = \frac{\hbar}{2} \vec{\sigma} incorporates the Pauli matrices for the , and \vec{L} is the orbital . This coupling lifts the degeneracy in the total quantum number j = l \pm 1/2, producing energy shifts proportional to (Z \alpha)^4 m c^2 / n^3, matching observed shifts and to high precision. The exact Dirac energy levels are given by E_{nj} = m c^2 \left[ 1 + \left( \frac{Z \alpha}{n - (j + 1/2) + \sqrt{(j + 1/2)^2 - (Z \alpha)^2}} \right)^2 \right]^{-1/2}, demonstrating how the Pauli matrices encode the contribution essential for this relativistic phenomenology. The Weyl representation of the Dirac matrices further emphasizes the role of Pauli matrices in describing chiral aspects of relativistic fermions, particularly in the massless limit relevant to high-energy processes. In this basis, \gamma^5 = i \gamma^0 \gamma^1 \gamma^2 \gamma^3 is block-diagonal as \gamma^5 = \begin{pmatrix} -I_2 & 0 \\ 0 & I_2 \end{pmatrix}, enabling simple chiral projections P_{L,R} = (1 \mp \gamma^5)/2 that separate left- and right-handed components. For massless Weyl fermions, the two-component spinors satisfy the i \partial_t \psi = \pm \vec{\sigma} \cdot \vec{p} \, \psi, and the helicity projectors are (1 \pm \vec{\sigma} \cdot \hat{p})/2, where \hat{p} = \vec{p}/|\vec{p}| projects onto states with definite along the direction. This structure, built directly from the Pauli matrices, underlies the chiral of weak interactions and the of neutrinos as Weyl fermions.

Quantum Information

In quantum information theory, the Pauli matrices serve as fundamental single-qubit , denoted as X = σ_x, Y = σ_y, and Z = σ_z, which implement bit-flip, bit-and-phase-flip, and phase-flip operations, respectively. The X gate swaps the computational basis states |0⟩ and |1⟩, corresponding to a 180-degree around the x-axis on the ; the Z gate introduces a relative phase of π between |0⟩ and |1⟩, acting as a 180-degree around the z-axis; and the Y gate combines a bit flip with a phase shift, equivalent to a 180-degree around the y-axis. These gates are Hermitian, unitary, and square to the identity, making them Clifford operations essential for universal quantum computation when combined with non-Clifford gates like the T gate. The Pauli group on a single qubit, consisting of elements {±I, ±X, ±Y, ±Z, ±iI, ±iX, ±iY, ±iZ}, forms an abelian group under multiplication up to phases and is central to the stabilizer formalism for quantum error correction. In this framework, a stabilizer code is defined by an abelian subgroup S of the n-qubit Pauli group (generated by independent Pauli operators) whose common +1 eigenspace encodes the logical qubits, enabling detection and correction of Pauli errors without disturbing the code space. The formalism, introduced by Gottesman, allows efficient simulation of Clifford circuits via the Gottesman-Knill theorem, as stabilizer states evolve under such operations while tracking only the group generators. This structure underpins measurement-based quantum computation and fault-tolerant architectures, where stabilizers project onto error syndromes. Pauli errors—bit flips (X), phase flips (Z), or both (Y)—are the primary noise models in quantum error correction codes like the Steane code and surface code. The [[7,1,3]] Steane code, a CSS code derived from the classical [7,4,3] Hamming code, encodes one logical qubit into seven physical qubits and corrects any single-qubit Pauli error by measuring six stabilizer generators (three X-type and three Z-type), identifying the error location via syndrome lookup tables. Similarly, the surface code, a topological stabilizer code on a 2D lattice, protects against Pauli errors through local plaquette (Z-stabilizers) and vertex (X-stabilizers) measurements, achieving a high error threshold (around 1%) for fault-tolerant scaling as the lattice size increases. In both codes, errors are decoded by minimum-weight matching of syndromes, ensuring logical fidelity improves exponentially with code distance for depolarizing noise. For gate synthesis in near-term quantum hardware, the Cartan decomposition of the single-qubit unitary group SU(2) facilitates efficient compilation, expressing any unitary as U = exp(i \mathbf{a} \cdot \boldsymbol{\sigma}) \exp(i b Z) \exp(i \mathbf{a} \cdot \boldsymbol{\sigma}), where \mathbf{a} = (a_x, a_y, a_z) parameterizes rotations around axes in the xy-plane for the outer terms, and the inner Z rotation handles phases. This form leverages native Z-rotations common in superconducting and trapped-ion devices, reducing pulse complexity compared to general Euler decompositions, with relevance amplified in 2020s hardware where arbitrary single-axis rotations are calibrated but multi-axis control is costly. The decomposition requires solving for parameters via numerical optimization or analytic formulas, enabling sub-microsecond gate times with fidelities exceeding 99.9%. Pauli matrices quantify entanglement in Bell states, the maximally entangled two-qubit states such as the singlet |Ψ^−⟩ = (|01⟩ - |10⟩)/√2, where the correlation function ⟨( \boldsymbol{\sigma} \cdot \mathbf{n} )_A \otimes ( \boldsymbol{\sigma} \cdot \mathbf{n} )_B ⟩ = -1 for any unit vector \mathbf{n}, violating classical Bell inequalities. These correlations, arising from the operator \boldsymbol{\sigma}_A \cdot \boldsymbol{\sigma}_B with expectation value -3 for the singlet, underpin protocols like quantum teleportation and dense coding, where local Pauli measurements extract nonlocal information. In multipartite extensions, such as GHZ states, Pauli correlations detect genuine multipartite entanglement beyond bipartite reductions.

Classical Analogues

In classical mechanics, the components of angular momentum \mathbf{L} satisfy the Poisson bracket algebra \{L_i, L_j\} = \epsilon_{ijk} L_k, which mirrors the Lie algebra \mathfrak{su}(2) of the Pauli matrices, where the commutation relations are [\sigma_i, \sigma_j] = 2i \epsilon_{ijk} \sigma_k. This structural analogy allows the normalized angular momentum \mathbf{L}/\hbar in phase space to serve as a classical counterpart to the Pauli matrices \boldsymbol{\sigma}, facilitating semiclassical approximations in quantization schemes for rotational systems. A prominent semi-classical application arises in the for a classical or gyroscope precessing in a slowly varying , where the accumulated is the solid angle subtended by the field's path on the unit in parameter space. This exhibits a monopole-like singularity at the origin, analogous to the Dirac monopole structure inherent in the SU(2) bundle geometry underlying the Pauli matrices, with the monopole strength corresponding to the moment's magnitude. Unlike the quantum Berry phase, which is quantized for half-integer , the classical version yields a continuous Hannay angle, yet retains the topological monopole feature from the \mathfrak{su}(2) algebra. In polarization optics, the Pauli matrices provide a natural basis for representing the state of polarized light via Jones vectors, which are two-component complex vectors analogous to states. The , describing the polarization ellipse, can be expressed as the expectation values \mathbf{S} = \langle \boldsymbol{\sigma} \rangle, where the Jones vector plays the role of the , enabling compact descriptions of birefringent and dichroic effects through SU(2)-like transformations on the . This formalism highlights the between light polarization and the two-level structure encoded by the Pauli matrices, without invoking . Rigid body dynamics offers another classical analogue through Euler's equations, which govern the evolution of \boldsymbol{\omega} on the \mathfrak{so}(3) \cong \mathfrak{su}(2) equipped with a left-invariant determined by the body's tensor. The equations take the form I \dot{\boldsymbol{\omega}} + \boldsymbol{\omega} \times (I \boldsymbol{\omega}) = 0, where the cross product reflects the \mathfrak{su}(2) bracket, and solutions trace geodesics on the rotation group SO(3) covered by SU(2). This geometric formulation underscores the role of Pauli-like generators in describing torque-free motion, linking classical rotations to the group's invariant structure.

References

  1. [1]
    11 More Two-State Systems - Feynman Lectures - Caltech
    (Anyone who is going to work in quantum physics really has to memorize them.) They are also called the Pauli spin matrices after the physicist who invented them ...
  2. [2]
    [PDF] Pauli Spin Matrices
    Thus, in order for operators to have the analogous behavior in matrix mechanics, operators must turn vectors into vectors. As it turns out this is the most ...
  3. [3]
    [PDF] Phys 506 lecture 1: Spin and Pauli matrices
    Any 2×2 matrix can be expressed in terms of the identity and the 3 Pauli matrices. This is called completeness over the space of 2 × 2 matrices. Check: α1 + βσx ...
  4. [4]
    [1308.2493] On quantum circuits employing roots of the Pauli matrices
    Aug 12, 2013 · The Pauli matrices are a set of three 2x2 complex Hermitian, unitary matrices. In this article, we investigate the relationships between certain ...
  5. [5]
    [PDF] Lecture 3 - USC Viterbi
    square. The frace of a square,. = Tr { A + B } ← matrix is the sum of the diagonal elements. The Pauli matrices are Hermitian, Unitary, to the identity, and ...
  6. [6]
    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.
  7. [7]
    [PDF] Problem Set 2 - UCLA Physics & Astronomy
    The Pauli matrices σ1, σ2, σ3 form a complete basis for all 2 × 2 Hermitian matrices. The Pauli matrices are: σ1 = ( 0 1. 1 0. ) , σ2 = ( 0 −i i 0. ) , σ3 ...
  8. [8]
    [PDF] LECTURE NOTES ON QUANTUM MECHANICS
    Note that the Pauli matrices have the properties, σ2 i = 1,. (1.41). {σi ... These matrices are Hermitian, traceless, and obey simple commutation relations,.
  9. [9]
    [PDF] Lecture Notes for Ph219/CS219: Quantum Information Chapter 2
    (By identifying J as the angular-momentum, we have implicitly chosen units with h = 1.) The Pauli matrices also have the properties of being mutually anticom-.
  10. [10]
    [PDF] On the quantum mechanics of magnetic electrons
    “ Zur Quantenmechanik des magnetischen Elektrons,” Zeit. f. Phys. 43 (1927), 601-623. On the quantum mechanics of magnetic electrons. By W. PAULI, Jr. in ...
  11. [11]
    [PDF] Introduction to Lie Groups and Lie Algebras Alexander Kirillov, Jr.
    This book is an introduction to the theory of Lie groups and Lie algebras, with emphasis on the theory of semisimple Lie algebras.
  12. [12]
    Pauli Two-Component Formalism - Richard Fitzpatrick
    It is also easily seen that the Pauli matrices satisfy the anti-commutation relations. $\displaystyle \{ \sigma_i, \sigma_j \} = 2 \,, (492). Here, $ \{a,b ...
  13. [13]
    [PDF] Theory of Angular Momentum and Spin
    The Pauli matrices satisfy very special commutation and anti-commutation relationships. ... The Pauli matrices obey the following anti-commutation properties.<|control11|><|separator|>
  14. [14]
    [PDF] Quantum Mechanics Charles B. Thorn1 - UF Physics
    where the braces signify an anticommutator instead of a commutator. ... orthogonal basis of the Lie algebra with the inner product defined as TrA†B ...
  15. [15]
    [PDF] Vectors and Matrices
    In this chapter, we present basic definitions and results of complex vectors and matrices needed in the study of quantum information science and quantum.
  16. [16]
    [PDF] HOMEWORK ASSIGNMENT 13: Solutions
    ... Sx, Sy, and Sz in the basis of eigenstates of Sy. A cyclic permutation (relabeling) of the indices leaves the commutation relations unchanged. Thus we can ...Missing: 1927 paper
  17. [17]
    [PDF] Physics 130C Lecture Notes, Winter 2014 Chapter 1: Quantum ...
    Mar 7, 2014 · More generally, it is often useful to have an explicit expression for the eigenvectors of n · ~σ in the σz eigenbasis. Here is a slick way ...
  18. [18]
  19. [19]
    [PDF] spin one-half, bras, kets, and operators - MIT OpenCourseWare
    Sep 17, 2013 · So each. Pauli matrix must have two eigenvalues that add up to zero. Since the eigenvalues can only be plus or minus one, we must have one of ...<|control11|><|separator|>
  20. [20]
    [PDF] Pauli Matrices: A Triple of Accardi Complementary Observables
    We show that the pair of Pauli matrices A, B associated to the unit directions α and β in R3 are Accardi complementary if and only if α and β are orthogonal if ...
  21. [21]
    None
    ### Summary: Pauli Matrices as Orthonormal Basis under Hilbert-Schmidt Inner Product
  22. [22]
    [PDF] At a very instrumental level, quantum theory is a set of rules ...
    This inner product is often called the Hilbert-Schmidt product. Show that the identity and the three Pauli matrices form an orthonormal basis, with respect to ...
  23. [23]
    [PDF] Solving the matrix exponential function for the groups SU(3 ... - arXiv
    Jul 8, 2022 · The well known analytical formula for SU(2) matrices U = exp(i~τ·~ϕ ) = cos |~ϕ| + i~τ· ˆϕ sin|~ϕ| is extended to the SU(3) group with eight ...
  24. [24]
    4.10 Matrix Exponentials - BOOKS
    ... Pauli matrices. For any such ... which shows explicitly how to relate the decompositions of Hermitian matrices and their corresponding unitary matrices.
  25. [25]
    [PDF] Lecture notes: Qubit representations and rotations
    Pauli matrices (18) satisfy the orthonormal- ity conditions (17) which is just the structure equation for the SU(2) group. [Si,Sj] = i ijk Sk, where Si ≡ σi. 2.
  26. [26]
    [PDF] Lecture notes Particle Physics II Quantum Chromo Dynamics 2. SU ...
    Nov 12, 2013 · These matrices are, together with the 2- or 3-dim vectors on which they act, called the fundamental representation of SU(2) or SU(3). • ...
  27. [27]
    [PDF] SU(2)'s Double-Covering of SO(3)
    Dec 12, 2019 · Thus, the generators of the Lie algebra of U(N) are the generators of the group of N × N Hermitian matrices and the identity I. Thus, the ZN ...<|control11|><|separator|>
  28. [28]
    [PDF] SU(2) to SO(3) homomorphism
    Simple connectedness is of great importance in Lie groups and Lie algebras because if a Lie group is simply connected, there is a one-to-one correspondence ...
  29. [29]
    [PDF] 0.1 Rotations of R3, the groups SO(3) and SU(2) - LPTHE
    Sep 9, 2012 · which is nothing else than the Rodrigues formula (1). We thus conclude that the transformation x 7→ x0 performed by the matrices of SU(2) in (14) ...
  30. [30]
    [PDF] Chapter 10 The Quaternions and the Spaces S , SU(2), SO(3), and RP
    tan θ. 2= #. N(t). |a|. , with 0 < θ ≤ π. We can write the unit quaternion Z as. Z = *cos θ. 2. , sin θ. 2V +, where V is the unit vector t. √N(t)(with −π ≤ θ ≤ ...
  31. [31]
    [PDF] Matrix Representation of Special Relativity - arXiv
    Mar 8, 2007 · The possibility of representing Minkowski spacetime vectors with 2x2-matrices has been known since the ... Let me start with the 4-dimensional ...
  32. [32]
  33. [33]
    [quant-ph/9705052] Stabilizer Codes and Quantum Error Correction
    Authors:Daniel Gottesman. View a PDF of the paper titled Stabilizer Codes and Quantum Error Correction, by Daniel Gottesman. View PDF. Abstract ...
  34. [34]
    [quant-ph/9605021] Simple Quantum Error Correcting Codes - arXiv
    May 15, 1996 · Access Paper: View a PDF of the paper titled Simple Quantum Error Correcting Codes, by Andrew Steane (Clarendon Laboratory and 1 other authors.
  35. [35]
    Poisson structure on a space with linear SU(2) fuzziness
    Jul 14, 2009 · For the group SU(2), the independent Poisson brackets (19)–(21) are in fact the Poisson structure of a rigid rotator, in which the angular ...
  36. [36]
    Classical adiabatic angle and geometrical phase in spin precession
    This Letter demonstrates the classical origin of the geometrical phase factor for a spin eigenstate which accompanies adiabatic traversal of a closed path, ...
  37. [37]
    Jones-matrix analysis with Pauli matrices: application to ellipsometry
    The coherency matrix formalism based on Pauli matrices is applied to analyze a general ellipsometer that is described by Jones matrices.
  38. [38]
    [PDF] Geometric Hydrodynamics in Open Problems
    Since the work of V. Arnold it is well-known that while the Euler equation of a rigid body corresponds to a left-invariant metric (depending on the body shape).<|control11|><|separator|>