Fact-checked by Grok 2 weeks ago

Tensor density

In , a tensor density (also known as a relative tensor) is a of a that transforms under a change of according to the standard tensor law multiplied by a factor of the of the matrix of the raised to a power called the weight w. The weight w is typically an (such as $0, \pm 1, \pm 2), though it can be any , and when w = 0, the object reduces to an tensor. Specifically, for a tensor density of type (k, l) with weight w, the components in the new coordinate system \overline{x} transform as \overline{T}^{j_1 \dots j_k}_{i_1 \dots i_l} = T^{s_1 \dots s_k}_{r_1 \dots r_l} \cdot \det\left( \frac{\partial x}{\partial \overline{x}} \right)^w \cdot \prod_{m=1}^k \frac{\partial \overline{x}^{j_m}}{\partial x^{s_m}} \cdot \prod_{n=1}^l \frac{\partial x^{r_n}}{\partial \overline{x}^{i_n}}, where x are the old coordinates and the products account for the usual tensor index transformations. Tensor densities arise naturally in contexts requiring coordinate-independent formulations of integrals, such as s on manifolds. For instance, the coordinate d^n x = dx^1 \wedge \dots \wedge dx^n in n-dimensional space is an n-form that behaves as a tensor density of -[1](/page/−1), transforming as d^n \overline{x} = \det\left( \frac{\partial \overline{x}}{\partial x} \right) d^n x to ensure the \int f \, d^n x is when f is a scalar. In Riemannian or pseudo-Riemannian , the scalar \sqrt{|\det g|}, where g is the , serves as a scalar density of +[1](/page/1), enabling the of the \sqrt{|\det g|} \, d^n x. These objects are particularly prominent in physics, especially in and field theories, where they facilitate the densitization of tensors for formulations or covariant integration. A classic example is the \varepsilon_{i_1 \dots i_n}, which is a tensor density of weight -1 rather than a true tensor, as its transformation includes an extra \det\left( \frac{\partial x}{\partial \overline{x}} \right)^{-1} factor; multiplying by \sqrt{|\det g|} yields the Levi-Civita tensor, which is a proper tensor. Tensor densities also appear in the on oriented manifolds, where they adjust for metric-induced volume scalings in . In non-orientable manifolds, modified versions like ψ-densities incorporate bundles to allow integration.

Motivation and Fundamentals

Role in Coordinate Transformations

In , standard tensors, which transform linearly under coordinate changes via the without additional scaling factors, fail to preserve volumes or integrals when subjected to diffeomorphisms that distort local metrics or orientations. For instance, the volume element defined by a standard tensor, such as the Cartesian dx dy dz, alters its measure under a nonlinear coordinate , leading to inconsistencies in physical or geometric quantities like or across manifolds. This limitation arises because diffeomorphisms can stretch or compress regions unevenly, requiring compensation to maintain invariance of integrals over arbitrary domains. The determinant addresses this by quantifying the local volume scaling under coordinate transformations, serving as the key factor in the transformation laws for tensor densities. Specifically, it measures the of the partial of the new coordinates with respect to the old ones, |det(∂x'/∂x)|, which multiplies the original to yield the transformed one, ensuring that densities adjust accordingly for uniform behavior. In this way, tensor densities incorporate powers of the to counteract the , allowing quantities like probability densities or densities to remain consistent across coordinate systems. Tensor densities thus function as essential tools for achieving invariant integration over manifolds, where integrals of density fields yield coordinate-independent results, such as total mass or in variational principles. By embedding the Jacobian scaling directly into their structure, these objects enable the construction of volume forms that are preserved under diffeomorphisms, facilitating computations in curved spaces or non-orthogonal coordinates without adjustments. This approach ensures that geometric invariants, like those in symplectic mechanics or , hold regardless of the chosen chart. The historical motivation for tensor densities traces back to early 20th-century differential geometry, particularly Élie Cartan's development of integral invariants in the 1920s, where densities were introduced to preserve measures under transformations in continuous media and variational problems. In his 1922 work, Cartan employed densities, such as mass per unit volume ρ, to form absolute integral invariants like ∫∫∫ ρ δx δy δz over volumes, addressing the failure of unadjusted forms to remain invariant under diffeomorphisms. This framework laid the groundwork for modern uses in invariant integration, emphasizing densities' role in handling orientation and volume preservation in geometric contexts.

Distinction from Standard Tensors

Standard tensors, whether contravariant or covariant, are multilinear maps that transform linearly under changes of coordinates via the partial derivatives of the coordinate functions, without any additional scaling factors. This ensures that their components in different coordinate systems are related in a way that preserves the intrinsic geometric structure, independent of the specific coordinate choice. In contrast, tensor densities incorporate a weight factor w, modifying the standard tensor transformation by multiplying it with the absolute value of the Jacobian determinant raised to the power w. Pseudotensors form a related but distinct category, typically referring to densities that also change sign under orientation-reversing transformations, such as the , which behaves as a pseudotensor density of weight -1. Physically, tensor densities describe quantities that scale with volume elements, such as density, which transforms with weight -n in n-dimensional to maintain invariance of total under coordinate rescaling. Similarly, probability densities in continuous distributions scale inversely with volume to preserve total probability. Tensor densities generalize standard tensors by accounting for the scaling introduced by the in non-orthogonal or curvilinear coordinate systems, enabling consistent descriptions of volume-dependent quantities.

Formal Definition

General Transformation Law

A tensor density of weight w generalizes the transformation behavior of standard tensors by incorporating an additional factor involving the of the coordinate . Under a change of coordinates from x^\mu to x'^\nu, the components of a scalar tensor density \phi transform according to \phi'(x') = \phi(x) \det \left( \frac{\partial x^\rho}{\partial x'^\sigma} \right)^w. This law accounts for the scaling under orientation-reversing transformations, where the sign of the affects densities with non-zero integer weights. For a general (k, l)-tensor density T^{\mu_1 \dots \mu_k}_{\nu_1 \dots \nu_l} of weight w, the transformation law extends the standard tensor rule by multiplying it with the factor: T'^{\alpha_1 \dots \alpha_k}_{\beta_1 \dots \beta_l}(x') = \left( \prod_{i=1}^k \frac{\partial x'^{\alpha_i}}{\partial x^{\mu_i}} \right) \left( \prod_{j=1}^l \frac{\partial x^{\nu_j}}{\partial x'^{\beta_j}} \right) T^{\mu_1 \dots \mu_k}_{\nu_1 \dots \nu_l}(x) \det \left( \frac{\partial x^\rho}{\partial x'^\sigma} \right)^w. The products of partial derivatives arise directly from the chain rule applied to the multilinearity of tensor components under . The factor, raised to the power w, is introduced to account for the scaling of the volume element d^n x' = \det \left( \frac{\partial x'}{\partial x} \right) d^n x, which modifies the transformation beyond the pure tensor case. This convention with the signed is for oriented manifolds, where the sign may vary, allowing densities to reflect .

Parity and Weight Classification

Tensor densities are classified by their w, which is typically a , most commonly an , that determines the power to which the is raised in the law. Specifically, under a coordinate with matrix J, the components transform as T' = (\det J)^w \cdot ( tensor ). When w = 0, this reduces to the transformation law for tensors, recovering their behavior without the additional factor. The of a tensor refers to its behavior under orientation-reversing coordinate transformations, where \det J < 0. Assuming |\det J| = 1 for simplicity (as in orthogonal transformations), the factor (\det J)^w = \operatorname{sign}(\det J)^w = (-1)^w. Thus, for even weights (w even), the is under such reversals, exhibiting even similar to true tensors but with by |\det J|^w. For odd weights (w odd), the acquires an extra minus sign, exhibiting odd akin to pseudotensors, which is essential for quantities sensitive to , such as oriented volumes. A representative example is the \varepsilon_{i_1 \dots i_n}, which serves as a covariant pseudotensor of weight -1 (odd) in n dimensions. It changes sign under orientation reversal, reflecting its role in defining oriented structures like determinants and volume elements. In contrast, the scalar \sqrt{|\det g|} has weight -1 (odd) and also flips sign, while products like (\sqrt{|\det g|})^2 = |\det g| have even weight +2 and remain invariant.
PropertyEven Weight (e.g., w = 0, \pm 2)Odd Weight (e.g., w = \pm 1)
Behavior under orientation reversal (\det J = -1) (factor +1)Changes sign (factor -1)
Parity typeEven (true tensor-like)Odd (-like)
Example of $\det g
ImplicationsSuitable for absolute volumes or Essential for oriented volumes and antisymmetric forms

Algebraic Properties

Operations and Contractions

Tensor densities support standard tensor algebraic operations, with modifications to account for their weight under coordinate transformations. These operations preserve the overall structure while adjusting the appropriately to maintain the density character. and follow rules analogous to those for ordinary tensors but are restricted by weight compatibility, while contractions and differentiations inherently preserve the . Addition of tensor densities is defined component-wise, but only between densities of the same type (i.e., same contravariant and covariant index ranks) and the same weight w. If T and S are tensor densities of weight w, their sum T + S is also a tensor density of weight w, transforming under coordinate changes as \overline{T + S} = | \det(\partial x / \partial \bar{x}) |^w ( \bar{T} + \bar{S} ). This ensures the Jacobian factor is identical for both terms, avoiding inconsistencies in the transformation law. Multiplication of tensor densities, typically via the , combines two densities of weights w_1 and w_2 to yield a new density of weight w_1 + w_2. For instance, if T is a density of weight w_1 and S of weight w_2, the product T \otimes S transforms as \overline{T \otimes S} = | \det(\partial x / \partial \bar{x}) |^{w_1 + w_2} (\bar{T} \otimes \bar{S}), reflecting the additive nature of the weights in the power. This operation increases the total by the of the individual ranks and is fundamental for constructing higher-order densities from simpler ones. Contraction, the summation over a pair of contravariant and covariant indices, reduces the of a tensor by two while preserving its w. For a mixed tensor T^i_j of w, the is given by C = T^i_i = \sum_{i=1}^n T^i_i, and the resulting scalar transforms as \bar{C} = | \det(\partial x / \partial \bar{x}) |^w \bar{T}^i_i, where the barred components incorporate the standard tensor transformation factors. This preservation arises because the sums terms that each carry the same factor, ensuring the overall remains unchanged. Partial differentiation commutes with the density structure, such that the partial derivative of a tensor density of weight w yields another tensor density of the same weight w. For a component T^{\mu_1 \dots \mu_k}_{\nu_1 \dots \nu_l}, the partial derivative \partial_\rho T^{\mu_1 \dots \mu_k}_{\nu_1 \dots \nu_l} transforms with the additional index \rho as a standard tensor component, multiplied by the overall | \det(\partial x / \partial \bar{x}) |^w factor, without altering the weight. This property holds because partial derivatives introduce no additional determinant factors beyond the original transformation law. The covariant derivative extends this, incorporating connection terms adjusted by -w \Gamma^\sigma_{\sigma \rho} to maintain the weight.

Special Cases for Weights Zero and One

Tensor densities of weight zero correspond precisely to ordinary tensors, as their transformation law under coordinate changes lacks any additional determinant factor. Specifically, a quantity T'^{\mu_1 \dots \mu_k}_{\nu_1 \dots \nu_l} = \frac{\partial x'^{\mu_1}}{\partial x^{\alpha_1}} \cdots \frac{\partial x^{\beta_l}}{\partial x'^{\nu_l}} T^{\alpha_1 \dots \alpha_k}_{\beta_1 \dots \beta_l} follows the standard multilinear transformation without modification, encompassing familiar objects such as vectors, covectors, and the . This equivalence ensures that weight-zero densities integrate over coordinate volumes without corrective factors, preserving tensorial consistency in applications like covariant derivatives. In contrast, tensor densities of one exhibit a distinctive role in defining volume measures, particularly scalar densities that facilitate over manifolds. A scalar density of one transforms as \phi' = \left| \det \frac{\partial x}{\partial x'} \right| \phi, incorporating the inverse to compensate for volume distortions under coordinate transformations. In , the factor \sqrt{|g|}, where g = \det(g_{\mu\nu}) is the of the , serves as a prototypical example of such a scalar density of one; it combines with the coordinate d^4x to yield the proper volume \sqrt{|g|} \, d^4x, essential for formulating principles and conservation laws. This property underscores their utility in ensuring diffeomorphism- integrals, such as those in the Einstein-Hilbert . Transitions between different weights in tensor densities can be achieved by multiplication with powers of a scalar density of unit weight, such as \sqrt{|\det g|}. To raise the weight from w to w+1, one multiplies by \sqrt{|\det g|}; conversely, lowering the weight involves division by \sqrt{|\det g|}, effectively adjusting the density character while preserving the underlying tensor structure. This operation is particularly useful in hybrid formulations where ordinary tensors must be adapted to density contexts, such as in numerical implementations of . A notable physical arises for scalar densities of weight +1, exemplified by probability density functions in and . Under a , a probability density f transforms as \bar{f} = |J|^{-1} f, where J = \det \left( \frac{\partial \bar{x}}{\partial x} \right), ensuring the \int f \, d^n x = 1 remains across coordinate systems. This weight-plus-one behavior aligns joint probability densities with symmetric covariant tensor densities, facilitating their use in manifold-based statistical models.

Advanced Manipulations

Inversion and Determinant Formulas

Tensor densities of weight w that are nonsingular as matrices possess inverses that are also tensor densities, but of weight -w. This follows from the requirement that the product of a density and its inverse yields the tensor, which has weight 0, necessitating the weights to sum to zero under the law. The components of the inverse are computed using the classical formula for matrix inversion: A^{-1} = \frac{1}{\det A} \adj A, where \adj A is the , obtained as the of the cofactor matrix of A. Under a coordinate with matrix J = \frac{\partial \overline{x}}{\partial x}, the inverse transforms according to the standard tensor rules for its type, multiplied by |\det J|^{-w}, ensuring consistency with the weight change. The determinant of a nonsingular tensor density matrix provides a scalar quantity whose transformation properties reflect both the tensor type and the density weight. For a (k,l)-tensor density of weight w in n-dimensional space (with k = l for a square matrix), the determinant scales under coordinate transformation as \det \overline{A} = \left| \det \left( \frac{\partial \overline{x}}{\partial x} \right) \right|^{k - l - w} \det A, where J = \frac{\partial \overline{x}}{\partial x} is the of the from old to new coordinates. This scaling arises because the determinant is a multilinear alternating form on the indices, with each of the k contravariant indices contributing a factor of \det J, each of the l covariant indices contributing a factor of (\det J)^{-1} through the , and the overall density factor (\det J)^{-w}. In particular, for balanced types like (1,1) with w = 0, the determinant is invariant, behaving as a scalar of weight 0. For the determinant of the g = \det g_{\mu\nu}, a (0,2)-tensor with w = 0, k=0, l=2, the exponent is $0 - 2 - 0 = -2, so \overline{g} = |\det J|^{-2} g, or equivalently g' = g \left[ \det \left( \frac{\partial x}{\partial \overline{x}} \right) \right]^2, consistent with standard calculations. To derive this, consider the multilinearity of the : \det A can be expressed as a over permutations involving products of components, each transforming linearly according to the tensor density . The full transformation of \det \overline{A} thus inherits the product of the individual index transformations, yielding (\det J)^k from contravariant indices, (\det J)^{-l} from covariant indices, and (\det J)^{-w} from the weight, for a total exponent k - l - w. This aligns with the known behavior of the of the itself, which for weight 1 transforms with exponent -1, confirming the structure for composite objects like densities. In linear algebra contexts, these formulas are essential for change-of-basis operations involving tensor densities, such as re-expressing multilinear forms or volume elements in new coordinates. For instance, when transforming a basis for a equipped with a density structure, the Jacobian accounts for the weight adjustment in the density, ensuring that traces or other contractions remain consistent across bases. This application is particularly useful in computational representations of , where inversion facilitates efficient coordinate shifts without altering the underlying geometric invariants.

Relation to Volume Forms

Tensor densities of weight 1 serve as a generalization of volume forms in differential geometry, extending the framework to non-oriented manifolds or situations without a predefined metric. On an oriented n-dimensional manifold, a volume form is a nowhere-vanishing n-form, such as dx^1 \wedge \cdots \wedge dx^n in local coordinates, which assigns an oriented volume element to each tangent space and enables coordinate-independent integration via pullbacks and partitions of unity. In contrast, a tensor density T of weight 1 transforms under coordinate changes via the factor \left| \det\left( \frac{\partial x}{\partial x'} \right) \right| or equivalently \left| \det\left( \frac{\partial x'}{\partial x} \right) \right|^{-1}, allowing it to define an unoriented volume element that scales appropriately with the Jacobian, thus providing a positive measure for integration without relying on an orientation. The conversion between a tensor density of weight 1 and a typically involves a or to produce a top-degree . For instance, on a , the volume form \omega = \sqrt{|\det g|} \, dx^1 \wedge \cdots \wedge dx^n arises by multiplying the coordinate (a density of weight +1) by the scalar density \sqrt{|\det g|} of weight -1, where g is the ; this ensures the form is a genuine tensor under diffeomorphisms. Similarly, in Lorentzian settings, \sqrt{-\det g} of weight -1 adjusts for the , yielding an for integrals like \int f \sqrt{-\det g} \, d^n x. This relation highlights how densities compensate for metric-induced scaling to yield orientable forms suitable for exterior calculus. The use of tensor densities ensures invariance of integrals under diffeomorphisms, as the weight-1 transformation law precisely counters the Jacobian factor from the coordinate change, making \int_M f T \, dx^1 \cdots dx^n well-defined and independent of the chart. This property is crucial for defining global measures on manifolds, such as in applications or submanifold integrations, where densities facilitate coordinate-free formulations. Positive definite densities of weight 1 are equivalent to volume elements on oriented manifolds, bridging the unoriented density bundle with the oriented top-form bundle. Extensions to fractional weights arise in specialized contexts like conformal geometry, where densities of non-integer weight w (e.g., w = 1/2) transform under conformal rescalings g' = \Omega^2 g as sections of line bundles \wedge^0, preserving conformal invariants. For oriented manifolds, the top-form bundle identifies with \wedge^0[-n], linking fractional densities to weighted volume measures in applications such as conformal hypersurface invariants or . These structures enable analysis in weighted or non-integer-dimensional settings, such as fractal boundaries or conformal field theories.

Applications in General Relativity

Integration with Metric Tensor

In , the integration of tensor densities with the g_{\mu\nu} is facilitated through the of the , denoted g = \det(g_{\mu\nu}). Under a coordinate from x to x', the components transform as g'_{\mu\nu} = \frac{\partial x^\alpha}{\partial x'^\mu} \frac{\partial x^\beta}{\partial x'^\nu} g_{\alpha\beta}, leading to the transforming as g' = \left[ \det\left( \frac{\partial x}{\partial x'} \right) \right]^2 g. Consequently, the quantity \sqrt{|g'|} = \left| \det\left( \frac{\partial x}{\partial x'} \right) \right| \sqrt{|g|}, which identifies \sqrt{|g|} as a scalar of weight +1, compensating for the factor in integrals to ensure invariance. In Lorentzian spacetimes with (-,+,+,+), this becomes \sqrt{-g}, where g < 0. The , being a tensor of weight 0, allows on tensor while preserving their weight. For a tensor T^{\mu_1 \cdots \mu_k}_{\nu_1 \cdots \nu_l} of weight w, contracting with g^{\rho\sigma} or g_{\rho\sigma} yields another of the same weight w, as the metric's transformation law does not introduce additional factors. This operation is essential for maintaining the character when manipulating indices in curved formulations. A key application is the invariant volume element in , given by dV = \sqrt{-g} \, d^4x, which transforms as a scalar under coordinate changes since \sqrt{-g'} \, d^4x' = \sqrt{-g} \, d^4x. This form ensures that integrals over scalar densities yield coordinate-independent results, crucial for action principles and conservation laws. Historically, following the finalization of the field equations in late 1915, Einstein incorporated \sqrt{-g} in his review to formulate invariant expressions for the gravitational action and field equations, enabling consistent handling of densities in generally covariant theories. This approach addressed the need for proper volume measures in curved , building on the developed with Grossmann.

Usage in Covariant Formulations

In , the covariant derivative of a tensor density T of weight w follows the standard tensorial form but incorporates an adjustment via the contracted to account for the density's transformation properties, without introducing additional weight-specific factors beyond these connection terms. Specifically, for a general tensor density, \nabla_\mu T^{\alpha_1 \cdots \alpha_p}_{\ \ \ \beta_1 \cdots \beta_q} = \partial_\mu T^{\alpha_1 \cdots \alpha_p}_{\ \ \ \beta_1 \cdots \beta_q} + \sum_{k=1}^p \Gamma^{\alpha_k}_{\mu \gamma} T^{\alpha_1 \cdots \gamma \cdots \alpha_p}_{\ \ \ \beta_1 \cdots \beta_q} - \sum_{l=1}^q \Gamma^\gamma_{\mu \beta_l} T^{\alpha_1 \cdots \alpha_p}_{\ \ \ \beta_1 \cdots \gamma \cdots \beta_q} - w \Gamma^\lambda_{\lambda \mu} T^{\alpha_1 \cdots \alpha_p}_{\ \ \ \beta_1 \cdots \beta_q}, where \Gamma^\lambda_{\nu\mu} are the of the second kind, and the \Gamma^\lambda_{\lambda \mu} = \frac{1}{\sqrt{|g|}} \partial_\mu \sqrt{|g|} links directly to the determinant g. This formulation ensures the result is itself a tensor density of the same weight, preserving the geometric consistency under coordinate changes. A practical manipulation involves converting tensor densities to true tensors for applying standard general relativity operations. By multiplying a density T of weight w by (\sqrt{|-g|})^{-w}, one obtains an ordinary tensor \tilde{T} = T (\sqrt{|-g|})^{-w} that transforms covariantly without the Jacobian factor. This allows operations like contraction, symmetrization, or index manipulation using the g_{\mu\nu}, after which the result can be converted back if needed by multiplying by (\sqrt{|-g|})^{w}. Such conversions are essential for maintaining invariance in calculations involving curved spacetimes. Tensor densities play a key role in action principles, where the Lagrangian density must be a scalar density of weight +1 to guarantee the invariance of the action integral under diffeomorphisms. For instance, in the Einstein-Hilbert action, the term \sqrt{-g} \, R combines the scalar curvature R (weight 0) with the metric density factor \sqrt{-g} (weight +1), yielding a density of weight +1 whose integral over coordinate volume elements is diffeomorphism-invariant. This property ensures the variational principle derives equations of motion that respect the general covariance of general relativity. In simulations, tensor densities provide significant advantages over purely coordinate-based approaches, which rely on partial derivatives and can introduce instabilities from coordinate artifacts. By employing densitized variables—such as the densitized lapse \tilde{\alpha} = \alpha \sqrt{\gamma} (where \gamma = \det \gamma_{ij} for the spatial ) or the conformal traceless extrinsic curvature in the BSSN —simulations maintain a covariant structure that enhances and accuracy in evolving the 3+1 of . These densities facilitate adaptive coordinate choices, reduce constraint violations, and avoid singularities in the evolution equations, enabling robust long-term simulations of phenomena like mergers.

Illustrative Examples

Cartesian to Curvilinear Coordinates

In two-dimensional Euclidean space, consider a scalar density \rho of weight w = -1, representing a coordinate area density such as the component for an invariant bare integral \int \rho \, d^2x, defined in Cartesian coordinates (x, y). Under a change to polar coordinates (r, \theta), where x = r \cos \theta and y = r \sin \theta, the Jacobian determinant of the transformation from polar to Cartesian coordinates is |\det(\partial(x,y)/\partial(r,\theta))| = r. The components of the scalar density transform according to the general law for tensor densities of weight w, \rho'(r, \theta) = \rho(x, y) \cdot |J|^{-w}, where J = |\det(\partial(r,\theta)/\partial(x,y))| = 1/r. For w = -1, this simplifies to \rho'(r, \theta) = \rho(r \cos \theta, r \sin \theta) / r. To compute this explicitly, first express \rho(x, y) in terms of polar variables: substitute x = r \cos \theta and y = r \sin \theta into the original density function. The adjustment factor $1/r accounts for the coordinate volume element in polar coordinates being dr \, d\theta without additional scale factors, ensuring the bare integral \int \rho' \, dr \, d\theta remains invariant. For instance, if \rho(x, y) = 1 (constant components in Cartesian), then \rho'(r, \theta) = 1/r. This transformation highlights how tensor densities differ from ordinary scalars, incorporating the local stretching of the coordinate grid. Note that this "uniform" refers to constant components in Cartesian; the corresponding physical density (with proper area element r \, dr \, d\theta) would vary as $1/r^2, reflecting the coordinate choice rather than physical uniformity. The invariant bare integral \int \rho \, dx \, dy = \int \rho' \, dr \, d\theta holds by construction for the density of weight -1, though care is needed near the due to the coordinate . For a vector density example, consider momentum density \mathbf{p} = \rho \mathbf{v}, a contravariant vector density of weight w = -1 analogous to charge current density in , where \rho is the scalar and \mathbf{v} is the velocity field. In Cartesian coordinates, the components are p^x(x, y) and p^y(x, y). In polar coordinates, the components transform as p'^r(r, \theta) = J^{-w} \left( \frac{\partial r}{\partial x} p^x + \frac{\partial r}{\partial y} p^y \right) and similarly for p'^\theta, with the weight adjustment J^{-w} = 1/r since J = 1/r and w = -1. The partial derivatives \partial r / \partial x = x/r = \cos \theta and \partial r / \partial y = \sin \theta provide the directional projection, while the $1/r factor scales the to maintain invariance under bare coordinate . If \mathbf{v} is radial, say \mathbf{v} = (x/r, y/r), the momentum density components adjust to preserve the total momentum flux in the invariant sense. This transformation ensures the integrated quantity over bare coordinates remains invariant, analogous to \int p^i \, d^2x being preserved. Visually, in polar coordinates, the density \rho' appears to decrease with radius for constant Cartesian components, reflecting the expanding circumferential "rings" that dilute the coordinate components while keeping the bare enclosed total constant, as illustrated by concentric circles where the coordinate element lacks the linear growth with r.

Stress-Energy Tensor in GR

In general relativity, the stress-energy tensor T^{\mu\nu} describes the distribution of , , and , serving as in the G^{\mu\nu} = 8\pi T^{\mu\nu}. This tensor is a contravariant of rank 2 and weight 0, meaning it transforms under general coordinate changes solely according to the standard tensor transformation law without additional factors. However, to formulate local conservation laws in a coordinate-independent manner, the combination \sqrt{-g}\, T^{\mu\nu} functions as a tensor of weight -1, where g = \det(g_{\mu\nu}) is the of the . The covariant conservation condition \nabla_\mu T^{\mu\nu} = 0 is equivalent to the coordinate expression \partial_\mu (\sqrt{-g}\, T^{\mu\nu}) = 0, ensuring that \sqrt{-g}\, T^{\mu\nu} behaves as a . This density structure is particularly evident when integrating to compute conserved quantities, such as the total energy-momentum across a spacelike hypersurface \Sigma. The 4-momentum component is given by P^\nu = \int_\Sigma T^{\mu\nu} n_\mu \sqrt{-g}\, d^3 x, where n_\mu is the unit normal to \Sigma; the factor \sqrt{-g}\, d^3 x provides the invariant proper volume element, rendering the integral scalar under coordinate transformations. In the Schwarzschild metric, which describes the spacetime around a spherically symmetric, non-rotating mass, the line element is ds^2 = -(1 - 2M/r) dt^2 + (1 - 2M/r)^{-1} dr^2 + r^2 d\Omega^2, yielding \sqrt{-g} = r^2 \sin\theta. Under a coordinate transformation, such as a radial rescaling r' = r + \delta r(r) with Jacobian determinant J = \det(\partial x^\alpha / \partial x'^\beta) \approx 1 + \partial_r \delta r for small perturbations, the transformed density \sqrt{-g'} T'^{\mu\nu} = (\sqrt{-g} T^{\mu\nu}) / |J| adjusts precisely to preserve the integral's value, as \sqrt{-g'} = \sqrt{-g} \, |J| from the metric's transformation law. This explicit Jacobian scaling highlights how tensor densities maintain physical invariance near coordinate-dependent features like the origin (r = 0), where the bare volume element vanishes. An illustrative application is the computation of the total within a static modeled in a curved akin to the Schwarzschild interior solution. For a stress-energy tensor T^{\mu\nu} = (\rho + p) u^\mu u^\nu + p g^{\mu\nu}, the integrated E = -\int T^0_0 \sqrt{-g}\, d^3 x over the stellar accounts for the local \rho (including rest and internal contributions) weighted by the proper . In coordinates where the exhibits singularities at the stellar (analogous to polar coordinate artifacts), the density form \sqrt{-g}\, T^{00} ensures the integral converges without artificial divergences, as the r^2 \sin\theta factor from \sqrt{-g} cancels the coordinate vanishing at r = 0, yielding a finite total consistent with the star's gravitational parameter M. For a constant-density of radius R, this yields E \approx 4\pi \int_0^R \rho r^2 dr, adjusted by relativistic factors near the boundary. In computational , tensor densities like \alpha \sqrt{\gamma} T^{\mu\nu} (with \alpha the lapse function and \gamma = \det(g_{ij}) the spatial ) are employed in 3+1 decompositions to formulate evolution equations in conservative form: \partial_t \mathbf{U} + \partial_i \mathbf{F}^i = \mathbf{S}, where \mathbf{U} includes densitized energy and momentum densities. This approach simplifies boundary conditions in numerical schemes, such as finite-volume methods, by treating fluxes naturally without explicit factors at boundaries, enhancing and accuracy in simulations of stellar collapse or binary mergers.

References

  1. [1]
    [PDF] an informal discussion on tensor calculus - CCoM
    Physics texts often introduce the concept of tensor density with a strange transfor- mation law and subsequently never give an invariant conception of them ( ...
  2. [2]
    Invariant Integrals and Tensor Densities
    A tensor density, , of weight transforms like a tensor except that the W th power of the Jacobian appears as a factor with the pattern shown below.
  3. [3]
    [PDF] Differential Geometry in Physics - UNCW
    slightly more complicated called a tensor density. Including the. √ det g is done in anticipation of this more general setting later. Since the forms dxi1 ...
  4. [4]
    [PDF] Differential Geometry in Physics
    Page 1. Differential. Geometry in. Physics. Gabriel Lugo. Differential Geometry ... tensor density of weight (−1). Instead, if g is any Riemannian or pseudo ...
  5. [5]
    [PDF] dr bob's elementary differential geometry - Villanova University
    Mar 4, 2018 · 3)-tensor density of weight −1. This tensor density has the form ijkωi ⊗ ωj ⊗ ωk in any basis, with numerically constant components, like ...
  6. [6]
    [PDF] Lessons on integral invariants - Neo-classical physics
    Page 1. ÉLIE CARTAN. Lessons on. Integral Invariants. Translated by. D.H. Delphenich. HERMANN. Editors for the sciences and arts, 156, boulevard Saint-Germain ...
  7. [7]
    2. Manifolds - Lecture Notes on General Relativity - S. Carroll
    The power to which the Jacobian is raised is known as the weight of the tensor density ... differential geometry: Stokes's theorem. This theorem is the ...
  8. [8]
    4.7: Things that Aren't Quite Tensors - Physics LibreTexts
    Mar 5, 2022 · In differential geometry, we do have a scalar product, which is ... tensors represent intensities, while tensor densities measure quantity.
  9. [9]
    [PDF] Differential Geometry
    In general, the covariant derivative of a tensor density need not be a tensor density. Another important example of a tensor density is the Levi–Civita alter-.
  10. [10]
    The Bundle of Tensor Densities and Its Covariant Derivatives - MDPI
    We construct the smooth vector bundle of tensor densities of arbitrary weight in a coordinate-independent way. We prove the general existence of a globally ...
  11. [11]
    [PDF] Tensors in computations
    The two goals are interwoven: what defines a tensor is also what makes it useful in computations, so it is important to gain a genuine understanding of tensors.
  12. [12]
    [PDF] INTRODUCTION TO VECTORS AND TENSORS - OAKTrust
    Volume II begins with a discussion of Euclidean Manifolds which leads to a development of the analytical and geometrical aspects of vector and tensor fields.<|control11|><|separator|>
  13. [13]
  14. [14]
    [PDF] arXiv:0911.0334v3 [gr-qc] 11 Jan 2024
    Jan 11, 2024 · Therefore, the covariant Levi-Civita symbol is a tensor density of weight -1 and its product with a scalar density is a tensor. The ...
  15. [15]
  16. [16]
    [PDF] Integration in General Relativity - arXiv
    We can convert tensor densities of weight w to ordinary tensors by noting ... which transforms like an ordinary tensor (i.e. tensor density of weight zero).
  17. [17]
    None
    ### Summary of Joint Probability Densities as Tensor Densities of Weight -1
  18. [18]
    density in nLab
    Jun 22, 2025 · A positive definite density is equivalently a volume element (or a volume form on an oriented manifold). 2. Definition. For X X a manifold its 1 ...
  19. [19]
    3 Introducing Riemannian Geometry‣ General Relativity by David ...
    The anti-symmetric tensor density arises in many places in physics. In all cases, it should be viewed as a volume form on the manifold. (In nearly all cases, ...
  20. [20]
    [PDF] Differential Forms - MIT Mathematics
    Feb 1, 2019 · One of the goals of this text on differential forms is to legitimize this interpretation of equa- tion (1) in 𝑛 dimensions and in fact, ...
  21. [21]
    [PDF] Conformal Loxodromes - arXiv
    Nov 19, 2023 · We may view conformal densities of weight 𝑤 as sections of a smooth line bundle ∧0 [𝑤] and observe that, if 𝑀 is oriented, then this ...
  22. [22]
    None
    Summary of each segment:
  23. [23]
    [PDF] Numerical Relativity: Solving Einstein's Equations
    ... tensor density rather than a tensor; see Appendix A.3. Page 70. 54. CHAPTER 3 ... densitized” lapse α = ψ6 ¯α,. (3.100) in terms of which (3.99) becomes.
  24. [24]
    [PDF] CHAPTER 4 GENERAL COORDINATES - Javier Rubio
    The absolute value in. Eq. (4.27) is introduced to take into account the case of a metric with Lorentzian signature (−+ ++) and negative determinant. A ...
  25. [25]
    [PDF] Slightly Bimetric Gravitation - arXiv
    Aug 16, 2007 · gravitational stress-energy tensor [38], not merely a pseudotensor ... A (1,1) tensor density of weight w is illustrative. For such a ...
  26. [26]
    None
    Summary of each segment:
  27. [27]
    [PDF] A non-relativistic limit of M-theory and 11-dimensional membrane ...
    Dec 21, 2023 · transforms as a rank two symmetric tensor of weight zero. It is ... Rollier, Boundary Stress-Energy Tensor and Newton-Cartan Geometry ...<|control11|><|separator|>