Fact-checked by Grok 2 weeks ago

Metric signature

In and linear algebra, the metric signature (or simply signature) of a g on a manifold refers to the of the associated , quantified by the ordered triple (p, q, r), where p is the number of positive eigenvalues, q the number of negative eigenvalues, and r the number of zero eigenvalues of the metric in a suitable basis. For non-degenerate metrics, r = 0, and the signature is often denoted as (p, q) or by the pattern of signs in the diagonalized form, such as (+, -, 0). This , protected by Sylvester's law of , classifies the geometry of the manifold, distinguishing (signature (n, 0)), (signature (1, n-1) or (n-1, 1)), and other pseudo-Riemannian structures. In the context of spacetime physics, particularly special and general relativity, the metric signature describes the causal structure of four-dimensional Minkowski spacetime, typically adopting a Lorentzian signature of (3, 1) or (1, 3) to reflect one timelike dimension and three spacelike dimensions. This choice ensures that the spacetime interval ds^2 = g_{\mu\nu} dx^\mu dx^\nu distinguishes timelike paths (for massive particles), spacelike separations, and null geodesics (for light), underpinning the light cone structure essential for causality and relativity principles. Two primary conventions prevail: the "mostly plus" signature (-, +, +, +), favored in general relativity texts where the time component is negative and proper time is \tau = \sqrt{-ds^2}, and the "mostly minus" signature (+, -, -, -), preferred in particle physics where the time component is positive and proper time is \tau = \sqrt{ds^2}. The signature profoundly influences formulations of physical theories, affecting signs in the , energy-momentum tensor, , and constraints, though the underlying physics remains invariant under global sign flips of the . In curved spacetimes, such as the Schwarzschild for black holes, the signature affects the —for example, in the (-, +, +, +) , the radial direction is spacelike outside the event horizon (r > 2M) and timelike inside (r < 2M), with the horizon itself being null—while in cosmology, it shapes the Friedmann–Lemaître–Robertson–Walker for universe expansion. Advanced research explores dynamical signature changes, potentially linking to quantum gravity or early universe models, but the standard Lorentzian signature remains foundational for describing our universe's geometry.

Fundamentals

Definition

In the context of real vector spaces, a bilinear form B: V \times V \to \mathbb{R} on a finite-dimensional vector space V is a function that is linear in each argument separately. For a symmetric bilinear form, where B(u, v) = B(v, u) for all u, v \in V, it can be represented in a basis by a symmetric matrix A, such that B(u, v) = u^T A v. This associated symmetric bilinear form motivates the quadratic form Q: V \to \mathbb{R} defined by Q(v) = B(v, v), which in matrix terms is Q(v) = v^T A v for v \in V. The metric signature of such a quadratic form Q is defined as the ordered triple (p, q, r), where p is the number of positive eigenvalues of A, q is the number of negative eigenvalues, and r is the multiplicity of the zero eigenvalue, with the dimension n = p + q + r. This signature is independent of the choice of basis due to . A quadratic form is non-degenerate if r = 0, meaning the associated bilinear form has trivial kernel, and in this case the signature simplifies to the pair (p, q) with p + q = n; otherwise, if r > 0, it is degenerate. Non-degenerate quadratic forms are central to defining metrics on spaces.

Notation and conventions

The metric signature of a quadratic form or bilinear form is commonly denoted by the ordered pair (p, q) in the non-degenerate case, where p represents the number of positive eigenvalues and q the number of negative eigenvalues of the associated symmetric matrix, with p + q = n equal to the dimension of the space. For the degenerate case, the notation extends to (p, q, r), where r is the multiplicity of the zero eigenvalue, or nullity, satisfying p + q + r = n. Alternative notations include the overall signature s = p - q, which captures the difference between positive and negative indices and is invariant under congruence, and the index \nu = q (or sometimes \nu = p), denoting the number of negative (or positive) eigenvalues. In mathematical literature, particularly in linear algebra and , the convention (p, q) consistently assigns p to the positive part and q to the negative part, with no requirement that p \geq q, though specific conventions may prioritize one form over the other in certain contexts like metrics. This arises from the of the into p terms of +1 and q terms of -1. In physics, notations emphasize the sign pattern of the diagonal , such as (+ , - , - , -) for Minkowski in the "mostly minus" convention (signature (1,3)) or (- , + , + , +) in the "mostly plus" convention (signature (3,1)), reflecting choices for the time component. The overall sign choice affects but not the underlying geometry, with "mostly plus" prevalent in , while "mostly minus" is common in and many texts. The notation evolved from James Joseph Sylvester's 1852 work on quadratic forms, where he introduced the "law of inertia" and described the invariants as the "number of positive squares," "number of negative squares," and "number of zero squares" in the canonical reduction, termed inertia indices—now denoted as (p, q, r)—with further developments influenced by and tensor analysis. Ambiguities in notation persist, particularly regarding whether p or q denotes the positive indices; however, the mathematical standard prioritizes p for positives, while physics contexts may reverse this for Lorentzian signatures to highlight the timelike dimension first (e.g., (1,3) for one positive time in mostly minus). In , the convention aligns with the mathematical usage, specifying (p, q) explicitly for pseudo-Riemannian metrics to avoid confusion in manifold classifications. The underlying structure ensures these notations remain basis-independent.

Properties

Dimension, rank, and signature

In the context of a quadratic form associated with a symmetric bilinear form on a vector space of dimension n, the parameters p, q, and r represent the numbers of positive, negative, and zero eigenvalues of the Gram matrix, respectively, satisfying the relation n = p + q + r. The rank \rho of the quadratic form, which equals the dimension of the image of the associated linear map, is given by \rho = p + q, corresponding to the number of non-zero eigenvalues. These quantities arise from the spectral decomposition over fields like the reals, where the eigenvalues determine the form's structure. For non-degenerate metrics, the vanishes, implying r = 0, so the simplifies to n = p + q and the equals the . In this case, the signature, often denoted by the pair (p, q), fully characterizes the metric up to , meaning two such metrics are equivalent under if and only if they share the same p and q. This notation (p, q, r) provides a for the indices, with the non-degenerate signature determining the class over the reals. The of the is formalized as the of the \rho into the positive part of size p and the negative part of size q, encapsulating the form's indefinite nature. While the of the equals the sum of all eigenvalues and the equals their product (up to sign conventions), these quantities do not uniquely determine the , as multiple (p, q) can yield the same and through varying eigenvalue magnitudes and signs. Thus, the provides essential additional invariants beyond scalar measures like and .

Sylvester's law of inertia

Sylvester's law of inertia asserts that for any real symmetric n \times n matrix A, there exists an invertible real matrix P such that P^T A P = \begin{pmatrix} I_p & 0 & 0 \\ 0 & -I_q & 0 \\ 0 & 0 & 0_r \end{pmatrix}, where I_k denotes the k \times k , p + q + r = n, p is the number of positive eigenvalues (counting multiplicities), q is the number of negative eigenvalues, and r is the multiplicity of the zero eigenvalue. This diagonal form is unique up to of the blocks, and the triple (p, q, r) is called the inertia of A. The theorem is named after , who provided a proof in 1852, building on earlier results by mathematicians such as Lagrange, Gauss, and Jacobi on the reduction of quadratic forms. Sylvester's contribution emphasized the invariance of the numbers of positive, negative, and zero terms in the under real linear transformations, drawing an analogy to the physical concept of as resistance to change. A standard proof proceeds by on the n. For n = 1, the result is immediate since A is a scalar that is positive, negative, or zero. Assuming the statement holds for dimension n-1, consider a nonzero entry in A, say a_{11} \neq 0. , assume a_{11} > 0 (the negative case is symmetric); complete the square for the corresponding to factor out a positive term, reducing to a smaller on the . Apply the induction hypothesis to this submatrix, then back-substitute to obtain the full diagonal form. Alternatively, the for symmetric matrices diagonalizes A = Q D Q^T with orthogonal Q and diagonal D, and scaling the columns of Q appropriately yields the desired congruence to the inertia matrix. The law implies that the signature (p, q) (or equivalently, the difference p - q) is an invariant of the under real transformations A \mapsto P^T A P for invertible P, independent of the choice of basis. This invariance ensures the existence of an for the in which the associated takes the canonical diagonal representation, facilitating the classification of quadratic forms up to .

Geometric interpretation

The metric signature (p, q, r) of a on a real vector space provides a geometric of the space into orthogonal corresponding to its positive, negative, and degenerate components, as enabled by the diagonalization guaranteed by Sylvester's law of inertia. The positive index p identifies a where the quadratic form is positive definite, resembling the standard Euclidean metric; here, the form measures squared lengths positively, and level sets such as \{ \mathbf{x} \mid Q(\mathbf{x}) = c > 0 \} trace ellipsoids, reflecting bounded, compact geometries akin to ordinary distances in \mathbb{R}^p. The negative index q corresponds to a subspace where the form is negative definite, contributing to an overall indefinite quadratic form when combined with the positive part; this introduces hyperbolic-like directions, where level sets for c \neq 0 form hyperboloids, unbounded in certain directions and allowing distinctions between regions where the form takes positive or negative values. In the two-dimensional case of signature (1,1), for instance, the equation x^2 - y^2 = 1 yields a , contrasting sharply with the ellipse of the positive definite (2,0) case x^2 + y^2 = 1, highlighting the saddle-like that mixes expansive and contractive behaviors. The zero index r defines a degenerate kernel, or radical, where the quadratic form vanishes identically, representing null directions with zero "length" under the metric; this subspace is isotropic, and the overall geometry collapses along these directions, leading to cylindrical or flat degeneracies in level sets that do not constrain motion within the kernel. In higher dimensions, such as signature (1, n-1, 0), the null set \{ \mathbf{x} \mid Q(\mathbf{x}) = 0 \} forms a cone-like structure separating the positive and negative regions, analogous to boundaries in pseudo-Euclidean spaces without imposing full definiteness.

Examples

Matrices and quadratic forms

The metric signature of a symmetric matrix is determined by the signature of the associated quadratic form, which counts the number of positive, negative, and zero eigenvalues in its diagonalized form under congruence. This signature, denoted (p, q, r) where p + q + r = n for an n × n matrix, classifies the form's definiteness and degeneracy. A simple example is the diagonal matrix \operatorname{diag}(1, 1, -1), which represents the quadratic form x^2 + y^2 - z^2. Its eigenvalues are 1, 1, and -1, yielding signature (2, 1, 0) with no zero eigenvalues, indicating a non-degenerate indefinite form. For a non-diagonal case, consider the 2 × 2 A = \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}, which corresponds to the quadratic form $2xy. This matrix is congruent to \operatorname{diag}(1, -1) via an appropriate , with eigenvalues 1 and -1, giving signature (1, 1, 0). A degenerate example is the diagonal matrix \operatorname{diag}(1, 0, -1), associated with the quadratic form x^2 - z^2. Here, the eigenvalues are 1, 0, and -1, resulting in signature (1, 1, 1), where the zero eigenvalue reflects the form's degeneracy along the y-direction. These matrix representations illustrate distinct metrics on \mathbb{R}^n: the Euclidean metric uses signature (n, 0, 0) for all positive eigenvalues, while the Lorentzian metric employs (1, n-1, 0) to capture one time-like and n-1 space-like directions. The indices p and q geometrically correspond to the dimensions of the maximal positive and negative definite subspaces.

Inner products in vector spaces

In finite-dimensional real vector spaces, a pseudo-inner product is defined as a B: V \times V \to \mathbb{R} that is non-degenerate and possesses a signature (p, q), where p is the number of positive eigenvalues and q the number of negative eigenvalues of the associated Q(v) = B(v, v), with p + q = \dim V for non-degenerate cases. This form generalizes the standard inner product by allowing indefinite s, where Q(v) can take positive, negative, or zero values depending on the direction of v. The V equipped with such a pseudo-inner product admits an orthogonal into subspaces corresponding to the positive, negative, and parts. Specifically, V = V_+ \oplus V_- \oplus V^\perp, where V_+ and V_- are the maximal subspaces on which the restriction of B is positive definite and negative definite, respectively, with dimensions p and q, and V^\perp is the (the of B), consisting of vectors orthogonal to all of V. For non-degenerate pseudo-inner products, the V^\perp = \{0\}, so V = V_+ \oplus V_-, and this is unique up to by Sylvester's of inertia. A representative example occurs on the space \mathbb{R}^{p+q} with the standard pseudo-inner product of signature (p, q), given by B(\mathbf{x}, \mathbf{y}) = \sum_{i=1}^p x_i y_i - \sum_{j=1}^q x_{p+j} y_{p+j}, where \mathbf{x} = (x_1, \dots, x_{p+q}) and \mathbf{y} = (y_1, \dots, y_{p+q}). This form is non-degenerate, as its matrix representation is the diagonal matrix \operatorname{diag}(I_p, -I_q) with full rank, and the associated quadratic form Q(\mathbf{x}) = B(\mathbf{x}, \mathbf{x}) yields the signature (p, q). In contrast, a positive definite inner product has signature (n, 0), inducing a \|\mathbf{v}\| = \sqrt{B(\mathbf{v}, \mathbf{v})} that satisfies the properties of a norm, whereas the indefinite case leads to a pseudo-norm \sqrt{|B(\mathbf{v}, \mathbf{v})|}, which fails to define a true since B(\mathbf{v}, \mathbf{v}) can be negative or zero for nonzero \mathbf{v}.

Computation

Methods for determining signature

The signature of a real A, which represents a Q(\mathbf{x}) = \mathbf{x}^T A \mathbf{x}, can be determined through the by computing the eigenvalues of A. By the , A is orthogonally diagonalizable as A = P \Lambda P^T, where \Lambda = \operatorname{diag}(\lambda_1, \dots, \lambda_n) contains the real eigenvalues \lambda_i, and the signature is the triple (n_+, n_-, n_0), with n_+ the number of positive \lambda_i, n_- the number of negative \lambda_i, and n_0 = n - n_+ - n_- the multiplicity of the zero eigenvalue. Conceptually, the eigenvalues can be found by solving the \det(A - \lambda I) = 0, or through iterative processes like the , which exploits the symmetry for efficient computation. An alternative theoretical approach uses congruence transformations to diagonalize A via row and column operations, akin to adapted for symmetric matrices. Starting with the , one performs elementary operations—such as or variable substitutions—to reduce it to a diagonal form \sum_{i=1}^n \delta_i y_i^2, where each \delta_i \in \{+1, -1, 0\}; the signature is then the counts of positive, negative, and zero entries on this diagonal. This process preserves the signature under Sylvester's law of , which guarantees that the number of sign changes is invariant across congruent matrices. The signature is formally given by the indices from this diagonal form, where the number of flips relative to an all-positive diagonal quantifies the negative eigenvalues under orthogonal transformations. However, these methods face limitations in near-degenerate cases, where small perturbations in A can cause eigenvalue crossings or ambiguities, leading to numerical instability in eigenvalue computations.

Numerical and algorithmic approaches

Computing the metric signature of a symmetric matrix numerically relies on robust linear algebra algorithms that approximate the inertia (p, q, r) while accounting for floating-point arithmetic limitations. Eigenvalue decomposition is a primary approach, where the eigenvalues of the matrix are computed via symmetric eigenvalue solvers such as those in the LAPACK library, which provides routines like DSYEVD for dense symmetric matrices. The signs of the eigenvalues directly yield the counts of positive (p), negative (q), and zero (r) contributions, but numerical precision issues arise when eigenvalues are near zero; thresholds (e.g., 10^{-12} times the matrix norm) are often applied to classify them definitively. An alternative method is the LDL^T decomposition, a variant of tailored for symmetric indefinite matrices, which factorizes the matrix as A = L D L^T with L lower triangular and D diagonal. During the process, the signs of the diagonal entries in D track the without full eigenvalue computation, enabling efficient determination by counting positive, negative, and zero pivots after symmetric pivoting for . This approach, implemented in libraries like LAPACK's DSYTRF routine, is particularly useful for large sparse matrices where full eigendecomposition is prohibitive. For a concrete workflow, consider a 3x3 such as A = [[0, 1, 0], [1, 0, 1], [0, 1, 0]]. First, apply LDL^T decomposition: perform symmetric elimination with row/column pivoting to obtain L and D, revealing one positive, one negative, and one zero diagonal entry (thus signature (1,1,1)). Verify by computing eigenvalues via a library routine, ensuring consistency within (e.g., using a of 10^{-10} for near-zero values). If the matrix is ill-conditioned ( > 10^6), preconditioning or iterative refinement may be needed to avoid misclassification of small eigenvalues. In software implementations, Python's library offers scipy.linalg.eigh for symmetric eigenvalue , allowing signature extraction by sign-counting the returned eigenvalues array, with built-in handling for near-zero values via the 'tol' parameter in related functions. Similarly, MATLAB's eig function on symmetric matrices provides eigenvalues for direct computation, and users can employ ldlt via third-party toolboxes or custom scripts to track pivot signs. For ill-conditioned cases, both environments recommend assessing the matrix with cond(A) and applying deflation techniques if eigenvalues cluster near zero.

Applications in Physics

Spacetime metrics in relativity

In special relativity, the geometry of spacetime is described by the Minkowski metric, which typically adopts the signature (1,3) or (+, −, −, −), where the time component is positive and the three spatial components are negative. This convention was implicitly chosen by in his 1905 paper on the electrodynamics of moving bodies, where the spacetime interval for light rays is expressed as x^2 + y^2 + z^2 - c^2 t^2 = 0, leading to the line element ds^2 = c^2 dt^2 - dx^2 - dy^2 - dz^2. An alternative convention, (3,1) or (−, +, +, +), places the negative sign on the time component and is prevalent in many general relativity texts, such as Misner, Thorne, and Wheeler's Gravitation, reflecting a "mostly plus" structure for spatial directions. The choice of signature has direct implications for classifying spacetime intervals between events. In the (+, −, −, −) convention, an interval is timelike if ds^2 > 0, meaning the events can be causally connected by a slower-than-light path; spacelike if ds^2 < 0, indicating no causal influence is possible; and null if ds^2 = 0, corresponding to light propagation. This (1,3) split enforces the light cone structure, where the future and past light cones bound the regions accessible to massive particles, preserving causality by distinguishing timelike directions (associated with proper time) from spacelike ones. In the (−, +, +, +) convention, the signs for timelike and spacelike intervals are reversed, but the physical interpretations remain equivalent after adjusting definitions. In general relativity, spacetime is modeled as a four-dimensional pseudo-Riemannian manifold equipped with a metric tensor of fixed Lorentzian , either (1,3) or (3,1), which ensures the theory's consistency with special relativity's causal structure on curved backgrounds. This fixed signature prevents acausal influences, as the metric's indefinite nature allows for timelike geodesics that define worldlines of observers while prohibiting closed timelike curves in standard solutions. Hawking and Ellis emphasize that the Lorentzian signature is essential for the global structure of spacetime, supporting theorems on singularity and causality in cosmological models. The adoption of these conventions traces back to Einstein's foundational work, but debates over the preferred sign have persisted in the literature due to varying emphases in particle physics (favoring +, −, −, − for positive energy-momentum) versus gravitational studies (often preferring −, +, +, + for positive spatial distances). These discussions, highlighted in comprehensive reviews of sign conventions, underscore that while the choice is arbitrary, consistency within a framework is crucial to avoid errors in curvature calculations and energy conditions.

Signature change in advanced theories

In advanced theories of quantum gravity, the metric signature can undergo a flip, interpreted as a phase transition from Lorentzian (3,1) to Euclidean (4,0) signatures, particularly in path integral formulations where Euclidean metrics facilitate convergence near singularities. This concept arises in the , where the wave function of the universe is computed via a path integral over compact Euclidean geometries that smoothly transition to Lorentzian spacetime, avoiding a sharp initial boundary and suppressing the . Such flips are motivated by the need to regularize quantum gravitational effects, with the Euclidean regime dominating at early times before reverting to Lorentzian as the universe expands. Examples of signature change appear in loop quantum gravity (LQG), where holonomy corrections at Planckian densities induce a dynamical transition from Lorentzian to Euclidean signature, effectively resolving cosmological singularities through a "bounce" mechanism. In loop quantum cosmology, a simplified LQG framework, this change occurs when the spacetime curvature reaches the Planck scale, modeled as an effective modification of the Hamiltonian constraint that alters the metric's causal structure. Similarly, in string theory, dualities such as T-duality and S-duality can map solutions between different signatures, effectively altering the number of timelike dimensions and linking Lorentzian type II string theories to Euclidean or mixed-signature supergravities, as seen in matrix model realizations. These transitions provide a theoretical bridge between perturbative string vacua and non-perturbative quantum gravity regimes. Mathematically, signature change poses challenges due to the discrete nature of possible signatures, precluding continuous transitions in classical general relativity; instead, it is modeled via analytic continuation through complex metrics or limits where the metric becomes degenerate with null directions (r > 0 in the signature (p, q, r)). In quantum settings, Wick rotations facilitate the flip, but ensuring continuity requires careful handling of the measure to avoid divergences, often via tunneling geometries that interpolate between signatures without violating Einstein's equations locally. As of 2025, ongoing research debates the implications of in causal dynamical triangulations (CDT), where numerical simulations reveal a spontaneous Wick-rotated phase at small scales, interpreted as a microscopic for the , potentially resolving divergences in . In asymptotic safety programs, formulations explore fixed-point behaviors that accommodate signature flips to maintain while achieving renormalizability, though approximations dominate computations. These developments lack experimental evidence but offer theoretical resolutions to singularities and the .

References

  1. [1]
    Metric Signature -- from Wolfram MathWorld
    The term metric signature refers to the signature of a metric tensor g=g_(ij) on a smooth manifold M, a tool which quantifies the numbers of positive, zero, ...
  2. [2]
    [PDF] General Relativity Fall 2018 Lecture 5: the metric tensor field
    Sep 21, 2018 · This number of pluses and minuses is called the signature of the metric. ... These coordinates are defined up to a Lorentz transformation, which, ...
  3. [3]
    [PDF] General Relativity, Black Holes, and Cosmology - JILA
    ... General Relativity. 51. 2.1. Motivation. 52. 2.2. The postulates of General ... The metric signature is −+++. Greek (brown) letters 𝜅, 𝜆, ..., denote ...
  4. [4]
    [PDF] Lecture 4.7. Bilinear and quadratic forms - Purdue Math
    Apr 9, 2020 · This means that every quadratic form has a well defined signature which is independent of its representation by a matrix. And every square ...
  5. [5]
    [PDF] introduction to the arithmetic theory of quadratic forms - Yale Math
    By definition, a quadratic form is a quadratic space with chosen coordinates. ... In particular, the bilinear form is symmetric if and only if its associated ...<|control11|><|separator|>
  6. [6]
    [PDF] Lectures on Differential Geometry Math 240C
    Jun 6, 2011 · pair (p, q) is called the signature of the pseudo-Riemannian metric. ... Christoffel symbols for the Lorentz metric (5.2) with index conventions 1 ...
  7. [7]
    [PDF] arXiv:1307.2171v2 [math.AT] 6 Aug 2013
    Aug 6, 2013 · This paper presents a generalisation of Sylvester's law of inertia to real non-degenerate quadratic forms on a fixed real vector bundle over a.
  8. [8]
    [PDF] Signatures in algebra, topology and dynamics
    As the title suggests, this paper reviews some classical and less classical properties of quadratic forms and their signatures. We tried to collect several ...
  9. [9]
    [PDF] Appendix E Metric convention conversion table - TU Darmstadt
    In this book we have systematically used the metric convention, ηµν = Diag[−1,+1,+1,+1], the “Pauli,” “East Coast,” or “mostly plus” metric.
  10. [10]
    [PDF] A demonstration of the theorem that every homogeneous quadratic
    ground of analogy, to give the name of the Law of Inertia for Quadratic. Forms, as expressing the fact of the existence of an invariable number inseparably ...
  11. [11]
    [PDF] arXiv:1211.4453v1 [math.DG] 19 Nov 2012
    Nov 19, 2012 · Let g be a pseudo-Riemannian metric on. M of signature (p, q). In the complex setting, we assume that J− is almost pseudo-. Hermitian (i.e. J−g ...
  12. [12]
    [PDF] Algebraic and Geometric Theory of Quadratic Forms
    ... Signature. 125. 34. Bilinear and Quadratic Forms Under Quadratic Extensions ... The even dimensional forms determine an ideal I(F) in the Witt ring of F ...
  13. [13]
    [PDF] notes on the algebraic theory of quadratic forms
    An n-ary quadratic form f also defines a quadratic map qf : kn → k given by evaluation of f. This function is quadratic in the sense that qf (λx) = λ2qf (x) ...
  14. [14]
    [PDF] Spectral theorems, SVD, and Quadratic forms
    8 Sylvester's law of inertia. Theorem (Sylvester's law of inertia). Let A be an n × n symmetric matrix, then there exists a unique pair of non-negative ...
  15. [15]
    [PDF] Quadratic form
    Mar 23, 2013 · In mathematics, a quadratic form is a homogeneous polynomial of degree two in a number of variables. For example, is a quadratic form in the ...
  16. [16]
    [PDF] Quadratic Forms and their Number Systems and Geometry - arXiv
    Apr 7, 2022 · A quadratic form is called degenerate or non- degenerate if the corresponding bilinear form is degenerate or nondegenerate. The vector space ...
  17. [17]
  18. [18]
    [PDF] Addendum on Quadratic Forms
    Dec 10, 2015 · Denoting the number of 1's in the above matrix by a+, the number of. −1's by a− and the number of 0's on the diagonal by a0, we call the triple ...
  19. [19]
    [PDF] BILINEAR FORMS The geometry of Rn is controlled algebraically by ...
    In particular, the trace, determinant, and (more generally) characteristic polynomial of a linear operator L: V → V are well-defined, independent of the choice ...
  20. [20]
    [PDF] Inner Product Spaces - Purdue Math
    Feb 16, 2007 · 11). The mapping (4.11.11) is called a pseudo-inner product in. R2 and, when generalized to R4, is of fundamental impor- tance in Einstein's ...
  21. [21]
    Sylvester's Inertia Law -- from Wolfram MathWorld
    Sylvester's Inertia Law: The numbers of eigenvalues that are positive, negative, or 0 do not change under a congruence transformation.
  22. [22]
    Computation of eigenvalues of a real, symmetric 3 × 3 matrix with ...
    Oct 28, 2022 · Here we seek to devise alternative formulas, equivalent but numerically stable, making use of only standard hardware and software tools to ...
  23. [23]
    [PDF] ON THE ELECTRODYNAMICS OF MOVING BODIES - Fourmilab
    This edition of Einstein's On the Electrodynamics of Moving Bodies is based on the English translation of his original 1905 German-language paper. (published as ...
  24. [24]
    [PDF] Gravitation - physicsgg
    This book follows the "Landau-Lifshitz Spacelike Convention" (LLSC). Arrows below mark signs that are "+" in it. The facing table shows signs that other authors.
  25. [25]
    1. Special Relativity and Flat Spacetime
    SPECIAL RELATIVITY AND FLAT SPACETIME. We will begin with a whirlwind tour of special relativity (SR) and life in flat spacetime.
  26. [26]
    Change of signature in classical relativity - IOPscience
    The authors point out that the classical Einstein field equations, suitably interpreted, allow a change of signature of spacetime. Specific examples of such ...
  27. [27]
    [PDF] Sign Conventions in General Relativity
    Dec 11, 2023 · General relativity has sign conventions for quantities like the metric tensor (mostly plus or minus) and curvature tensor. Some conventions are ...
  28. [28]
    [1207.4657] Signature change in loop quantum cosmology - arXiv
    Jul 19, 2012 · We show that the signature change anticipated by Hartle and Hawking naturally appear in loop quantum cosmology.
  29. [29]
    [hep-th/9807127] Duality and the Signature of Space-Time - arXiv
    Jul 17, 1998 · These theories are all linked by duality transformations which can change the number of time dimensions as well as the number of space ...
  30. [30]
    Signature change in matrix model solutions | Phys. Rev. D
    Oct 12, 2018 · Signature change is believed to be a feature of quantum gravity [1–10] . It has been discussed in the context of string theory [6] , loop ...
  31. [31]
    Alternative route towards the change of metric signature | Phys. Rev. D
    Sep 23, 2019 · While studying the tunneling solutions, it immediately became clear that it is impossible to transition a solution of the Einstein equations ...
  32. [32]
    [2501.03752] Foliated Asymptotically Safe Gravity: Lorentzian ... - arXiv
    Jan 7, 2025 · In this work, we use the Arnowitt-Deser-Misner (ADM) decomposition of the metric degrees of freedom and investigate the relations between the ...