Fact-checked by Grok 2 weeks ago

Symmetric tensor

In , a is a or an array of components that remains invariant under the of its indices, meaning that swapping any two indices does not alter its value. For a second-rank tensor A, this symmetry condition is expressed as A^{mn} = A^{nm}, distinguishing it from antisymmetric tensors where components change sign under such exchanges. Symmetric tensors generalize the notion of symmetric matrices, which are second-order cases where the matrix equals its own , to higher-order tensors that arise in and tensor analysis. Any tensor can be uniquely decomposed into the sum of a symmetric part and an antisymmetric part, with the symmetric component given by \frac{[1](/page/1)}{2}(T_{ab} + T_{ba}) for rank-two cases, allowing for the of symmetric structures in physical and geometric applications. In three dimensions, a symmetric rank-two tensor has six independent components and can be further split into a (scalar) part and a traceless part, facilitating representations in contexts like stress tensors in . Key properties of symmetric tensors include their ability to be diagonalized in orthonormal bases for rank-two cases, where they possess real eigenvalues and orthogonal eigenvectors, and their decomposition into sums of rank-one symmetric terms like outer products v \otimes \cdots \otimes v for higher ranks, known as the symmetric tensor . These features underpin applications in fields such as , blind source identification, and , where the minimal number of such rank-one terms (symmetric ) determines and . The product of a symmetric tensor and an antisymmetric one always vanishes, highlighting their orthogonal roles in tensor spaces.

Definition and Basics

Definition

In , a tensor of type (k,0) is a from the of k copies of a V to the underlying field, such as the real or complex numbers. A tensor T of type (k,0) over a V is symmetric if it is invariant under any of its arguments, meaning T(v_1, \dots, v_k) = T(v_{\sigma(1)}, \dots, v_{\sigma(k)}) for all vectors v_1, \dots, v_k \in V and any permutation \sigma in the S_k. In coordinates with respect to a basis of V, the components T_{i_1 \dots i_k} of a satisfy T_{i_1 \dots i_k} = T_{j_1 \dots j_k} whenever (j_1, \dots, j_k) is any permutation of (i_1, \dots, i_k). For mixed tensors of type (k,l), can be defined separately with respect to the k contravariant indices or the l covariant indices, requiring invariance under permutations within each group of like indices; full across all indices is possible but less common unless specified.

Examples

Symmetric tensors appear in various mathematical and physical contexts, providing concrete illustrations of their defining property of invariance under index permutation. For instance, a rank-0 tensor, which is simply a scalar such as or , possesses no indices and is thus trivially symmetric, as there are no permutations to consider. A rank-1 tensor, equivalent to a , is likewise always symmetric because it has only a single index, leaving no pairs available for swapping while preserving the tensor's value under transformations. In , the g_{ij} serves as a fundamental covariant example of a rank-2 , satisfying g_{ij} = g_{ji} to define distances and angles consistently on a manifold. This ensures the functions as a non-degenerate inner product on tangent spaces. Contravariantly, the tensor I^{ij} in exemplifies , with I^{ij} = I^{ji}, arising from its construction as I^{ij} = \sum_\alpha m_\alpha (\delta^{ij} r_\alpha^2 - x_\alpha^i x_\alpha^j), where the coordinate products enforce equality under index exchange due to rotational invariance. For higher ranks, a simple example of a fully symmetric rank-3 tensor is the outer product v \otimes v \otimes v for a vector v, which remains unchanged under any permutation of its three indices.

Properties and Operations

Basic Properties

A symmetric tensor of order k on a V is characterized by its invariance under the action of the S_k, which permutes the tensor's indices or arguments. Specifically, the space of symmetric k-tensors, denoted S_k(V), is the image of the symmetrizer operator P = \frac{1}{k!} \sum_{\sigma \in S_k} \sigma, where \sigma acts by permuting the factors in the tensor product V^{\otimes k}. This operator is a projection onto S_k(V), satisfying P^2 = P, and ensures that any element T \in S_k(V) remains unchanged under any : \sigma(T) = T for all \sigma \in S_k. This invariance defines the of symmetric tensors and distinguishes them from general multilinear maps. For symmetric bilinear forms, which correspond to order-2 symmetric tensors in S_2(V^*), the symmetry implies that the associated form \langle u, v \rangle = \langle v, u \rangle for all u, v \in V. Over the real numbers, if this form is additionally positive definite—meaning \langle x, x \rangle > 0 for all nonzero x \in V—it defines an inner product on V, endowing the space with a structure. This preservation of symmetry under argument swap is fundamental to the geometry induced by such tensors. The trace operation on a symmetric (0,2)-tensor T, defined as \operatorname{tr}(T) = T^i_i in a chosen basis (summing over repeated indices), is well-defined and independent of the basis due to the tensor's . This yields a scalar under coordinate changes, reflecting the tensor's intrinsic properties. Furthermore, orthogonal transformations preserve the symmetry of a tensor: if T is symmetric in one , its components remain symmetric after any orthogonal , ensuring consistency across rotated frames. In , the space S_k(V) forms an of the general linear group \operatorname{GL}(V), meaning it contains no proper nontrivial invariant subspaces under the natural action of \operatorname{GL}(V). This irreducibility underscores the fundamental role of symmetric tensors in decomposing tensor spaces and highlights their from other symmetry types. The of S_k(V) for \dim V = n is \binom{n+k-1}{k}, quantifying the in such representations.

Symmetric Part of a Tensor

The symmetric part of a tensor is obtained by applying the symmetrization , which projects a general tensor onto the of symmetric tensors by averaging over all permutations of its indices. For a rank-k tensor T \in V^{\otimes k} over a vector space V of dimension n, the symmetrization operator S: V^{\otimes k} \to V^{\otimes k} is defined as S(T) = \frac{1}{k!} \sum_{\sigma \in S_k} \sigma(T), where S_k is the symmetric group on k elements, and \sigma(T) denotes the action of the permutation \sigma by rearranging the tensor factors according to \sigma. This operator possesses key properties that make it a projection onto the symmetric subspace. Specifically, it is idempotent, satisfying S^2 = S, meaning applying S twice yields the same result as applying it once, and it commutes with the action of permutations, ensuring consistency under index relabeling. The image of S, denoted S_k(V), consists of all fully symmetric rank-k tensors, and its dimension is given by the binomial coefficient \binom{n+k-1}{k}, which counts the number of independent components. For mixed tensors, which have both contravariant and covariant indices, symmetrization can be applied partially over specific index sets rather than all indices. For instance, one may symmetrize only over a of contravariant indices by averaging over permutations within that group, while leaving other indices unchanged; the generalizes to S(T) = \frac{1}{m!} \sum_{\sigma \in S_m} \sigma(T) for an m-index , preserving the tensor's overall type. A concrete example illustrates this for a (0,2)-tensor A_{ij} in n dimensions: its symmetric part is (A_{ij} + A_{ji})/2, which averages over the two permutations of the covariant indices and reduces the independent components from n^2 to \binom{n+1}{2} = n(n+1)/2. Such operations preserve scalar invariants like the , as permutations do not alter the contraction along diagonal elements.

Symmetric Product

The symmetric product of two vectors v and w in a V over a F is defined as v \odot w = \frac{1}{2}(v \otimes w + w \otimes v), which lies in the space of symmetric bilinear forms or, equivalently, the second symmetric power S_2(V). This construction ensures that the result is invariant under interchange of v and w, yielding a symmetric rank-2 tensor. For higher ranks, the symmetric product generalizes through iterated applications: for vectors v_1, \dots, v_k \in V, the k-fold symmetric product is given by the symmetrizer \frac{1}{k!} \sum_{\sigma \in [S_k](/page/Symmetric_group)} v_{\sigma(1)} \otimes \cdots \otimes v_{\sigma(k)}, where S_k is the symmetric group on k elements, projecting into the k-th symmetric power S_k(V). The collection of all symmetric powers \bigoplus_{k=0}^\infty S_k(V) forms the S(V), which is a graded under the extension of the symmetric product, with multiplication induced by the followed by symmetrization. In coordinates with respect to a basis of V, the basis elements of S(V) are monomials x_1^{a_1} \cdots x_n^{a_n} where \sum a_i = k for the -k component, reflecting the commutative structure where the product satisfies v \odot w = w \odot v. Symmetric tensors in S_k(V) are in natural correspondence with homogeneous polynomials of degree k on the dual space V^*, via the map that sends a symmetric tensor T to the polynomial p_T(u^*) = T(u^*, \dots, u^*) for u^* \in V^*, or more generally, p_T(u_1^*, \dots, u_k^*) = T(u_1^*, \dots, u_k^*) as a symmetric multilinear form. For example, a quadratic symmetric tensor corresponds to a quadratic form q(v) = T(v, v), which defines a bilinear form on V. This duality underpins the universal property of S_k(V): any symmetric k-linear map from V^k to another space extends uniquely to a linear map from S_k(V). The Veronese embedding provides a geometric realization of symmetric tensors, mapping the projective space \mathbb{P}(V) to \mathbb{P}(S_k(V)) via \nu_k() = [v^{\odot k}], where v^{\odot k} = v \odot \cdots \odot v (k times), embedding into a projective variety whose points correspond to rank-1 symmetric tensors. This map, of degree k, highlights the role of symmetric products in , parametrizing hypersurfaces defined by homogeneous polynomials associated to higher-rank symmetric tensors.

Decompositions and Representations

Polar Decomposition

The polar decomposition theorem states that any invertible second-order tensor F, representing a linear map between vector spaces, can be uniquely factored as F = R U, where R is an orthogonal tensor satisfying R^T R = I and U is a positive definite symmetric tensor. This decomposition separates the rotational and stretching components of the transformation, with U capturing the pure deformation. The right stretch tensor U is explicitly given by U = \sqrt{F^T F}, which is always symmetric and positive definite for invertible F. A left polar form exists as F = V R, where V = \sqrt{F F^T} is the left stretch tensor, also positive definite and symmetric, while maintaining the same orthogonal R. Both forms are unique for invertible F, ensuring a canonical separation into orthogonal and symmetric positive definite factors. This applies specifically to (1,1)-tensors, or matrices in finite dimensions. For higher-rank tensors, the generalizes through higher-order forms that extend the matrix case via Kronecker-structured covariance matrices, decomposing a tensor X \in \mathbb{R}^{p_1 \times \cdots \times p_K \times n} into factors involving positive definite symmetric components across modes and an orthogonal core. In this extension, the symmetric stretch is captured by lower triangular matrices L_k with positive diagonals, forming positive definite P_k = L_k L_k^T, while the orthogonal part satisfies mode-wise conditions. Uniqueness holds under conditions such as n \geq p, though practical computations often succeed more broadly. Computation for the second-order case proceeds via () of F = Q \Sigma P^T, yielding U = P \Sigma P^T (with singular values as eigenvalues of U) and R = Q P^T. For higher ranks, an iterative block coordinate descent algorithm updates each mode's factor by applying matrix polar decomposition to unfolded slices, orthogonalizing the core simultaneously for stability. The eigenvalues of the symmetric stretch can be obtained via of the relevant Gram tensors.

Spectral Decomposition

The spectral theorem for real symmetric tensors provides a diagonalization in an orthonormal basis, analogous to the matrix case. For a real symmetric (1,1)-tensor, which corresponds to a T \in \mathbb{R}^{n \times n}, the theorem states that T can be decomposed as T = Q D Q^T, where Q is an whose columns are the eigenvectors of T, and D is a containing the real eigenvalues \lambda_1, \dots, \lambda_n of T. This decomposition arises because symmetric matrices have only real eigenvalues, and their eigenvectors can be chosen to form an , ensuring diagonalizability with no nontrivial Jordan blocks due to the symmetry. For higher-order symmetric tensors, the spectral decomposition generalizes to a multi-linear form using symmetric eigentensors. A symmetric tensor T \in S_d(\mathbb{R}^n) of order d \geq 3 admits a decomposition T = \sum_{i=1}^r \lambda_i u_i^{\otimes d}, where the \lambda_i are real eigenvalues associated with unit vectors u_i that are orthogonal (i.e., u_i^T u_j = \delta_{ij}) when the tensor is orthogonally decomposable, and r \leq n is the tensor rank. This extends the rank-2 case, with eigenvalues defined via the equation T u^{d-1} = \lambda u^{d-1} for H-eigenvalues (real eigenvectors) or T u^{d-1} = \lambda u for E-eigenvalues (normalized), both of which are real and invariant under orthogonal transformations. Multiplicity is handled by the number of distinct eigentensors, and the symmetry ensures orthogonality among them without complex Jordan-like structures. Computationally, for large symmetric tensors, the power iteration method iteratively approximates the dominant eigenvalue and eigenvector by updating u^{(k+1)} = T (u^{(k)})^{d-1} / \|T (u^{(k)})^{d-1}\|, converging linearly under suitable conditions like distinct leading eigenvalues. For rank-2 cases, the Lanczos algorithm efficiently computes extremal eigenvalues by tridiagonalizing the matrix through orthogonal transformations, offering faster convergence for sparse structures.

Young Tableau Representation

In , the structure of symmetric tensors is captured through Young diagrams, which classify the irreducible representations of both the S_k and the general \mathrm{GL}(V) for a V of n. For fully symmetric tensors of k, the corresponding Young diagram is the single-row partition $$, consisting of k boxes in one row; this labels the irreducible \mathrm{GL}(V)-representation S^{} V, which is precisely the k-th symmetric power \Sym^k V. The symmetric within the tensor V^{\otimes k} is obtained via the Young symmetrizer projector for the partition $$, defined as P_{} = \frac{1}{k!} \sum_{\sigma \in S_k} \sigma, where \sigma acts by permuting the tensor factors. This operator is the average over all elements of S_k and projects onto the under the action, yielding the fully symmetric tensors; unlike antisymmetrizers, it involves no sign characters since $$ corresponds to the trivial of S_k. Schur-Weyl duality provides the full decomposition of the tensor power under the commuting actions of \mathrm{GL}(V) and S_k: V^{\otimes k} \cong \bigoplus_{\lambda \vdash k} S^\lambda V \otimes U^\lambda, where the direct sum runs over all partitions \lambda of k, S^\lambda V is the irreducible \mathrm{GL}(V)-module (Schur module) labeled by \lambda, and U^\lambda is the corresponding irreducible S_k-module (Specht module). The symmetric tensors form the summand for \lambda = , where S^{} V is the symmetric subspace tensored with the one-dimensional trivial S_k-representation. The of this symmetric is \dim(S^{} V) = \binom{n + k - 1}{k}, which counts the number of monomials of degree k in n variables and arises as a special case of the Weyl dimension for Schur modules. For general partitions \lambda \vdash k, the of S^\lambda V is given by the hook-content : \dim(S^\lambda V) = \prod_{(i,j) \in \lambda} \frac{n + c(i,j)}{h(i,j)}, where the product is over boxes in the Young diagram of \lambda, c(i,j) = j - i is the content of box (i,j), and h(i,j) is its hook length (the number of boxes to the right and below, plus one for the box itself). In the symmetric case \lambda = , each of the k boxes in the single row has hook length k - j + 1 for the j-th box, reducing the formula to the \binom{n + k - 1}{k}; the underlying hook-length f^\lambda = k! / \prod h(i,j) originally computes the of the S_k-irreducible U^\lambda.

Applications

In Physics

In physics, symmetric tensors play a crucial role in describing laws and physical observables across various domains, from to and . Their symmetry often arises from fundamental symmetries of the underlying physical laws, ensuring consistency with principles like . The -energy tensor T^{\mu\nu}, which encodes the distribution of energy, momentum, and in , is symmetric (T^{\mu\nu} = T^{\nu\mu}) due to the of derived from applied to Lorentz invariance. This symmetry implies that the total , including contributions from orbital and parts, is conserved in isolated systems, a cornerstone of relativistic field theories. In the absence of , the canonical -energy tensor derived from the is automatically symmetric; for fields with intrinsic , a symmetrized version (e.g., Belinfante-Rosenfeld tensor) restores this property to match observations. In , the inertia tensor I_{ij} for a is a symmetric second-rank tensor that quantifies rotational about the center of , given by I_{ij} = \int (r^2 \delta_{ij} - x_i x_j) \, dm, where r is the distance from the axis, \delta_{ij} is the , and the integral is over the distribution. This symmetry follows from the scalar nature of the and ensures that the \mathbf{L} = \mathbf{I} \cdot \boldsymbol{\omega} and \boldsymbol{\tau} = \mathbf{I} \cdot \dot{\boldsymbol{\omega}} equations are well-defined for , simplifying the Euler equations for . In , the strength tensor F_{\mu\nu} is antisymmetric (F_{\mu\nu} = -F_{\nu\mu}), capturing the duality between electric and magnetic fields, whereas the associated electromagnetic stress-energy tensor remains symmetric to align with the general requirements of energy-momentum . This contrast highlights how properties distinguish field strengths from their energy-momentum contributions, with the symmetric stress-energy tensor ensuring compatibility with Noether-derived laws. In , the \rho represents the statistical state of a quantum system as a symmetric (Hermitian) that is with unity (\operatorname{Tr}(\rho) = 1), allowing computation of expectation values \langle A \rangle = \operatorname{Tr}(\rho A) for observables A. This form generalizes pure states (\rho = |\psi\rangle\langle\psi|) to mixed states, preserving probabilistic interpretations and enabling descriptions of open quantum systems and entanglement. In , the G_{\mu\nu}, defined as G_{\mu\nu} = R_{\mu\nu} - \frac{1}{2} R g_{\mu\nu}, is symmetric due to the , which enforce the covariant \nabla^\mu G_{\mu\nu} = 0. This ensures that the G_{\mu\nu} = \frac{8\pi G}{c^4} T_{\mu\nu} couple consistently to the symmetric stress-energy tensor, upholding energy-momentum in curved . The g_{\mu\nu}, also symmetric, provides the background for these relations.

In Differential Geometry

In differential geometry, symmetric tensors play a fundamental role on Riemannian manifolds, where they often arise as sections of tensor bundles that encode geometric structures invariant under certain symmetries. A key example is the metric tensor, which defines the geometry of the manifold. On a smooth manifold M, a Riemannian metric g is a smooth section of the bundle of symmetric bilinear forms on the tangent bundle, specifically a (0,2)-tensor that is symmetric (g(X,Y) = g(Y,X) for vector fields X, Y) and positive definite at each point, thereby inducing an inner product on each tangent space T_p M. This symmetry ensures that the metric preserves the inner product structure under parallel transport along geodesics, facilitating the measurement of lengths, angles, and volumes on the manifold. The , uniquely determined by the , introduces symmetric tensors through its connection coefficients, known as . These symbols \Gamma^k_{ij} are derived from the via the formula \Gamma^k_{ij} = \frac{1}{2} g^{kl} \left( \partial_i g_{jl} + \partial_j g_{il} - \partial_l g_{ij} \right), where g^{kl} is the inverse , and \partial denotes partial in local coordinates. The symmetry of the g_{ij} = g_{ji} implies that the are symmetric in their lower indices, \Gamma^k_{ij} = \Gamma^k_{ji}, which corresponds to the torsion-free property of the . This symmetry is crucial for defining covariant derivatives that are compatible with the , enabling the extension of to tensor fields while preserving the inner product. Another important symmetric tensor is the , which captures second-order geometric information. For a smooth f: M \to \mathbb{R} on a , the \nabla^2 f is defined as the second , \nabla^2 f (X, Y) = X(Y f) - (\nabla_X Y) f, and it forms a symmetric (0,2)- acting as a on vectors. The symmetry \nabla^2 f (X, Y) = \nabla^2 f (Y, X) follows from the torsion-freeness of the and the compatibility, making the a natural tool for studying critical points, geodesics, and variations. In local coordinates, its components are H_{ij} = \partial_i \partial_j f - \Gamma^k_{ij} \partial_k f, highlighting its role in the of submanifolds and optimization on manifolds. The , a conformally component of the , also exhibits involving symmetric parts in its index structure. Defined as the trace-free part of the , W^a_{bcd} = R^a_{bcd} - \frac{1}{n-2} (R^a_c g_{bd} - R^a_d g_{bc}) + \frac{R}{(n-1)(n-2)} (g^a_c g_{bd} - g^a_d g_{bc}) in n-dimensions, the Weyl tensor inherits the Riemann tensor's symmetries: antisymmetry in the last two indices (W^a_{bcd} = -W^a_{bdc}), antisymmetry in the first pair when raised (W_{abcd} = -W_{bacd}), and symmetry under pair exchange (W_{abcd} = W_{cdab}). These symmetries, including the symmetric interchange of index pairs, ensure that the Weyl tensor measures the intrinsic conformal distortion of the manifold, independent of local scaling. Conformal transformations further underscore the role of symmetric tensors by preserving their symmetry properties up to . A conformal change of is given by \tilde{[g](/page/G)} = e^{2\sigma} [g](/page/G), where \sigma is a smooth function on M, transforming the original symmetric [g](/page/G) to another symmetric \tilde{[g](/page/G)} since the exponential factor is a scalar multiple that commutes with the . This rescaling leaves invariant while altering lengths, and the remains unchanged under such transformations, highlighting how symmetric tensors like the maintain their structural integrity in conformal geometry.

References

  1. [1]
    Symmetric Tensor -- from Wolfram MathWorld
    A second-tensor rank symmetric tensor is defined as a tensor A for which A^(mn)=A^(nm). Any tensor can be written as a sum of symmetric and antisymmetric parts.
  2. [2]
    [PDF] 6 Tensors
    Another way of saying this is that any symmetric rank 2 tensor has a basis of orthonor- mal eigenvectors {ei}, with Ii the corresponding eigenvalues. In the ...<|control11|><|separator|>
  3. [3]
    [PDF] SYMMETRIC TENSORS AND SYMMETRIC TENSOR RANK - HAL
    Abstract. A symmetric tensor is a higher order generalization of a symmetric matrix. In this paper, we study various properties of symmetric tensors in ...
  4. [4]
    [PDF] Multilinear Algebra and Tensor Symmetries
    Aug 28, 2011 · Let f : V ×···× V → W (k factors) be any k-multilinear map that is symmetric in its arguments. There is a unique linear map F : Sk(V) → W such ...
  5. [5]
    [PDF] exploiting symmetry in tensors for high performance
    row, we denote it as bai. 2.2. Symmetric tensors. A symmetric tensor is a tensor that is invariant under any permutation of indices. Definition 1 (Symmetric ...<|control11|><|separator|>
  6. [6]
    [PDF] Introduction to Tensor Calculus
    It is not useful to speak of symmetry or anti-symmetry in a pair of indices that are not of the same type (co- or contravariant). The properties of symmetry ...
  7. [7]
    [PDF] 7 Tensors - DAMTP
    A tensor of rank 0 is just a number, or scalar, T. Under a rotation, it doesn't change: T/ = T. A tensor of rank 1 is a vector, while a tensor of ...
  8. [8]
    [PDF] Lecture Notes for PHY 405 Classical Mechanics
    Nov 25, 2003 · Also note that Ii,j = Ij,i ⇒ the inertia tensor is symmetric. The ... The principal moments of inertia are thus I1 = I2 = (m1+m2)b2 ...
  9. [9]
    Lecture Notes on General Relativity - S. Carroll
    The Riemann tensor, with four indices, naively has n4 independent components in an n-dimensional space. In fact the antisymmetry property (3.64) means that ...
  10. [10]
    bilinear form in nLab
    ### Summary of Symmetric Bilinear Forms and Related Properties
  11. [11]
    [PDF] Physics 114A Tensors Peter Young
    There are also quantities of importance in physics which have more than one index and transform into each other in a more complicated way, to be defined below.
  12. [12]
    [PDF] Group Theory - Birdtracks
    The trace of the symmetrization operator yields the number of independent compo- nents of fully symmetric tensors: dS = tr S = ... =n + p − 1 p ... =(n + p ...
  13. [13]
    [PDF] arXiv:2203.01809v1 [math.AP] 3 Mar 2022
    Mar 3, 2022 · Given indices i1, ··· ,im the operator of partial symmetrization with respect to the indices i1,...,ip, where p<m, of a tensor u ∈ Tm is given ...
  14. [14]
    Section 10.13 (00DM): Tensor algebra—The Stacks project
    We define the symmetric algebra \text{Sym}(M) of M over R to be the quotient of \text{T}(M) by the two sided ideal generated by the elements x \otimes y - y \ ...
  15. [15]
    [PDF] MULTILINEAR ALGEBRA 1. Tensor and Symmetric Algebra Let k be ...
    By a k-algebra, we mean a ring homomorphism φ : k → A such that each element of Imφ commutes with each element of A. (If A is a ring, we define its center to be ...
  16. [16]
    Polar Decomposition | SpringerLink
    Apr 1, 2018 · The polar decomposition theorem states that any deformation gradient tensor can be multiplicatively decomposed into the product of an orthogonal tensor.
  17. [17]
    [PDF] A higher-order LQ decomposition for separable covariance models
    Oct 4, 2014 · We develop a higher order generalization of the LQ decomposition and show that this de- composition plays an important role in ...
  18. [18]
    [PDF] Spectral theorems, SVD, and Quadratic forms
    Theorem (Spectral theorem, version II). Let A be an n × n symmetric matrix, then all eigenvalues of A are real, their algebraic multiplicities and geometric ...
  19. [19]
    [PDF] Orthogonal Decomposition of Symmetric Tensors - arXiv
    Jun 17, 2015 · Abstract. A real symmetric tensor is orthogonally decomposable (or odeco) if it can be written as a linear combination of symmetric powers ...
  20. [20]
    Eigenvalues of a real supersymmetric tensor - ScienceDirect.com
    In this paper, we define the symmetric hyperdeterminant, eigenvalues and E-eigenvalues of a real supersymmetric tensor.
  21. [21]
    [PDF] The Spectral Theory of Tensors (Rough Version) - arXiv
    Jan 17, 2012 · The spectral theory of tensors is an important part of numerical multi-linear algebra, or tensor computation [48, 76, 90].
  22. [22]
    Subspace power method for symmetric tensor decomposition - arXiv
    Dec 9, 2019 · We introduce the Subspace Power Method (SPM) for calculating the CP decomposition of low-rank real symmetric tensors.
  23. [23]
    (PDF) The Lanczos Algorithm for Solving Symmetric Linear Systems
    The Lanczos Algorithm is becoming accepted as a powerful tool for finding the eigenvalues and eigenvectors of large sparse matrices.
  24. [24]
    Young Tableaux - Cambridge University Press & Assessment
    The aim of this book is to develop the combinatorics of Young tableaux and to show them in action in the algebra of symmetric functions, representations of ...
  25. [25]
    [PDF] Schur-Weyl Duality - UChicago Math
    We will compute the Young symmetrizers for all partitions of 3 and, thus, all the distinct irreducible representations of S3. Here are all the canonical tableau ...
  26. [26]
    [PDF] Noether's Theorems and Energy in General Relativity - arXiv
    Mar 31, 2021 · An important property of the quasi-local energy-momentum tensors is that they can lead to covariant integral conservation laws under weaker ...
  27. [27]
    [PDF] 3D Rigid Body Dynamics: The Inertia Tensor - MIT OpenCourseWare
    For an axisymmetric body, the moments of inertia about the two axis in the plane will be equal. Therefore, the moment about any axis in this plane is equal to ...Missing: source | Show results with:source
  28. [28]
    [PDF] 5. Electromagnetism and Relativity - DAMTP
    J = 0, (5.45) tells us that energy in the electromagnetic field is conserved. ... is known as the stress-energy tensor, or sometimes as the energy-momentum tensor ...
  29. [29]
    [PDF] Density Matrices and Quantum Operations
    A density matrix is a positive semi-definite Hermitian operator with Tr ρ = 1 (see remarks below). A2′ The mean value of an observable a is given by. ⟨A⟩ = Tr ( ...Missing: symmetric | Show results with:symmetric
  30. [30]
    [PDF] Revisiting the Bianchi identity 12.2 Making a theory of gravity - MIT
    The quantity in parenthesis is the as-yet-unnamed tensor Gαµ which we briefly noted in the previous lecture as the “trace-reversed” Ricci curvature.
  31. [31]
    [PDF] Introduction to Tensor Calculus for General Relativity - MIT
    The metric tensor has only. 10 independent coefficients (because it is symmetric). From equation (74), we see that we are left with 6 degrees of freedom for ...
  32. [32]
    The spectral geometry of the Weyl conformal tensor - math - arXiv
    Oct 15, 2003 · Abstract: We study when the Jacobi operator associated to the Weyl conformal curvature tensor has constant eigenvalues on the bundle of unit ...