Fact-checked by Grok 2 weeks ago

Einstein notation

Einstein notation, also known as the , is a notational in and physics that implies over a set of indices that are repeated within a term of an expression, thereby simplifying the representation of operations on vectors, matrices, and tensors without explicit summation symbols. This stipulates that an index appearing exactly twice in a single term—once as a subscript and once as a superscript, or both as subscripts in Cartesian coordinates—is to be summed over its entire range, unless explicitly stated otherwise. For instance, the expression a_i b_i denotes \sum_i a_i b_i, where i is a dummy index. The notation was introduced by in 1916 as part of his development of the general , where it facilitated the compact expression of complex tensor equations central to describing curvature and gravitational effects. In his seminal paper, Einstein described the rule as follows: "If an index occurs twice in one term of an expression, it is always to be summed unless the contrary is expressly stated," a convention he later humorously credited with suppressing cumbersome signs to streamline his work. Prior to this, had been formalized by and in the late 19th and early 20th centuries, but Einstein's adoption and popularization of the implicit made it indispensable for relativistic physics. Einstein notation distinguishes between free indices, which appear only once per term and define the components of the resulting tensor, and dummy indices, which are repeated and thus summed over, allowing for elegant manipulation of multi-dimensional quantities. It is particularly useful in operations like contractions, where tensors are multiplied and summed over shared indices, as in the scalar product c = a_i b^i or the determinant involving the Levi-Civita symbol \epsilon_{ijk}. The convention extends naturally to contravariant (upper indices) and covariant (lower indices) tensors, accommodating the for in curved spaces, and it integrates seamlessly with the \delta_{ij} for identity operations and the permutation symbol for cross products. Beyond general relativity, Einstein notation has become a standard tool in quantum field theory, continuum mechanics, and engineering applications involving multi-axial stresses, where it reduces notational clutter and highlights structural symmetries in equations. Its computational advantages are evident in modern implementations, such as preprocessor macros in programming languages that automate index handling for numerical simulations. Despite its efficiency, the notation requires careful adherence to rules—such as ensuring no index appears more than twice per term—to avoid ambiguity, making it a powerful yet precise shorthand for advanced mathematical physics.

Introduction

Core Convention

Einstein notation, also known as the Einstein summation convention, is a notational used in to imply over repeated indices in an expression, eliminating the need for explicit symbols. In this convention, when an index appears exactly twice in a term—once as a subscript and once as a superscript, or both as subscripts in Cartesian coordinates—it is understood to be summed over all possible values from 1 to the dimension of the , typically denoted as n. This approach streamlines the of multivariable equations in fields such as physics and , where tensors with multiple indices are . A fundamental example illustrates this implicit summation: the expression a_i b^i represents \sum_{i=1}^n a_i b^i, where the repeated index i (lower in a_i and upper in b^i) indicates the sum without writing the \sum symbol. Here, lower indices denote covariant components, while upper indices denote contravariant components, distinguishing the tensorial nature of the quantities involved. The rule requires that repeated indices appear once in each position (subscript and superscript) to trigger summation; an index appearing more than twice or only once does not imply summation and must be handled explicitly if needed. The convention is attributed to , who introduced it in his 1916 paper on to simplify the notation for complex tensor equations in curved . By convention, this pairing of upper and lower indices aligns with the metric tensor's role in , ensuring consistency in relativistic calculations. One key advantage of Einstein notation is its ability to simplify multi-index equations, reducing visual clutter from signs and allowing for more intuitive manipulation of tensor products and contractions. This compactness not only enhances readability but also facilitates algebraic proofs and computational implementations, making it particularly valuable in and .

Historical Development

The development of tensor calculus, which laid the groundwork for index-based notations, began in the late 19th century with the work of Gregorio Ricci-Curbastro and Tullio Levi-Civita. Ricci introduced key concepts of absolute differential calculus as early as 1887, evolving them through publications in the 1890s, while Levi-Civita collaborated on refining these ideas into a systematic framework by 1900. Their seminal 1900 paper employed index notation for tensors but relied on explicit summation symbols (Σ) for repeated indices, lacking a standardized convention for implicit summation. Albert Einstein popularized the implicit summation convention in 1916 while formulating the field equations of general relativity, introducing it explicitly in his foundational paper to streamline the manipulation of complex tensor expressions. He stated that "if an index occurs twice in one term of an expression, it is always to be summed unless the contrary is expressly stated," thereby transforming the cumbersome explicit sums of prior tensor calculus into a more concise form without inventing the underlying index system. This convention, now known as Einstein notation, was not original to him but marked its first prominent use in physics literature, facilitating the clarity of relativity's mathematical structure. Following Einstein's adoption, the notation gained standardization in the 1920s through the efforts of mathematicians like Jan Arnoldus Schouten, who integrated and formalized it in his works on tensor analysis and unified field theories, such as his 1922 book Der Ricci-Kalkül. By the mid-1920s, it appeared routinely in differential geometry and physics textbooks, and by the 1930s, it had become a standard tool in the field, bridging the Italian school of Ricci-Levi-Civita with broader applications in relativity. In the mid-20th century, Einstein notation became ubiquitous in and , enabling compact representations of multilinear operations in curved spacetimes and particle interactions. Its evolution accelerated in the late 20th century with the rise of computer algebra systems; for instance, Mathematica incorporated support for it via the EinsteinSummation function in 2019, while SymPy's tensor module has implemented the convention since around 2011.

Fundamentals of the Notation

Index Conventions and Summation Rule

In Einstein notation, indices are placed as superscripts or subscripts to distinguish between contravariant and covariant components of tensors. Upper indices denote contravariant components, such as v^i, which transform inversely to the basis vectors, while lower indices denote covariant components, such as v_i, which transform directly with the dual basis or covectors. This placement reflects the choice of basis: contravariant indices align with the basis, and covariant with the basis, ensuring consistent transformation properties under coordinate changes. The rule, central to the notation, implies that a repeated —appearing exactly twice in a term, once as an upper and once as a lower in general coordinates, or both as lower indices in Cartesian coordinates—is to be summed over its entire range, without the need for an explicit summation symbol. No summation occurs if an appears only once (a free ) or more than twice, the latter being a notational error that invalidates the expression. This convention, introduced by Einstein to streamline tensor equations in , applies strictly to such paired indices to avoid ambiguity in multi- expressions. For multi-index tensors, the rule facilitates compact representation of operations like matrix multiplication. Consider the product of two second-rank tensors forming another, expressed as c^i_j = a^i_k b^k_j, where the repeated index k (upper in a^i_k and lower in b^k_j) implies summation over k, yielding the i-th row and j-th column element of c. The range of this summation typically spans from 1 to n, the dimension of the space; in the context of special or general relativity, it often runs from 0 to 3 for spacetime coordinates. Common errors in applying these conventions include incorrectly raising or lowering an index without a metric tensor, or mismatched pairings of upper and lower indices that fail to form valid repeated pairs, leading to nonsensical or dimensionally inconsistent results. Such mistakes disrupt the tensorial nature of expressions and must be avoided to maintain the notation's rigor. A variant known as abstract index notation uses placeholder letters, such as T^a_b, to indicate the type and valence of a tensor without specifying numerical components or a particular basis, emphasizing abstract tensorial structure over concrete indices. This approach, developed by Penrose, leverages the same summation rule but treats indices as labels for tensor slots rather than summation variables.

Free and Dummy Indices

In Einstein notation, indices are classified as either free or dummy based on their role within an expression or equation. A free index appears exactly once on each side of an equation and determines the rank and type (or valence) of the tensor it describes; for instance, in the mixed tensor T^i_j, the indices i (upper) and j (lower) are free, indicating a rank-2 tensor with one contravariant and one covariant index. These free indices must match in both number and position (upper or lower) between the left and right sides of any valid equation to ensure the equality holds for the tensor's structure and preserves its type during manipulations. This matching rule guarantees that tensor equations remain consistent, as mismatched free indices would imply an imbalance in the tensor's dimensionality or variance. Dummy indices, in contrast, are those that appear twice in a single term—once raised and once lowered—and imply an implicit summation over all possible values of that index, as per the Einstein summation convention. Unlike free indices, dummy indices serve as placeholders for the summation process and can be freely relabeled without altering the equation's meaning; for example, the expression a_i b^i = c_j d^j equates two independent sums (over i and over j), and renaming the dummy index i to m on the left side yields a_m b^m = c_j d^j, which remains equivalent. In multi-term equations, dummy indices need not match between terms or sides, provided the free indices do, allowing flexible manipulation while the summation is preserved. A concrete illustration is the matrix multiplication A^i_j = B^i_k C^k_j, where i and j are free indices defining the output tensor's structure, and k is the dummy index over which summation occurs. These distinctions have key implications for verifying tensor equality and performing algebraic operations, as they enforce type preservation and prevent errors in index tracking. In practice, notation variants often distinguish index types by alphabet: Latin letters (e.g., i, j, k) typically denote spatial indices running from 1 to 3, while Greek letters (e.g., \mu, \nu) are reserved for spacetime indices running from 0 to 3 in relativistic contexts, aiding clarity in distinguishing summation ranges.

Representations in Linear Algebra

Vectors and Covariant/Contravariant Forms

In Einstein notation, vectors are represented as linear combinations of basis vectors, with components distinguished by their index position to reflect their transformation properties under coordinate changes. A contravariant \mathbf{v} is expressed as \mathbf{v} = v^i \mathbf{e}_i, where v^i are the contravariant components with upper indices, and \mathbf{e}_i denotes the basis vectors of the . This notation emphasizes that contravariant components transform in a manner that preserves the vector's direction relative to the coordinate axes. Covariant vectors, or covectors, are elements of the and are represented as \mathbf{w} = w_i \boldsymbol{\epsilon}^i, where w_i are the covariant components with lower indices, and \boldsymbol{\epsilon}^i forms the dual basis satisfying \boldsymbol{\epsilon}^i (\mathbf{e}_j) = \delta^i_j. These components arise naturally in contexts like gradients, where the covector acts on a contravariant to produce a scalar via the Einstein summation convention. In spaces equipped with an , the distinction between upper and lower indices is often omitted, allowing a unified subscript notation such as v_i for components, since the identifies contravariant and covariant forms (v_i = v^i). This simplification is common in basic linear algebra applications, where the basis eliminates the need for index positioning to track transformations. To aid intuition, contravariant vectors are sometimes mnemonicized as "column vectors" due to their role in describing positions or inputs, while covariant vectors correspond to "row vectors" that operate on them to yield scalars. Under a coordinate transformation from x^j to x'^i, contravariant components transform as v'^i = \frac{\partial x'^i}{\partial x^j} v^j, reflecting their scaling with the new coordinates, whereas covariant components transform inversely as w'_i = \frac{\partial x^j}{\partial x'^i} w_j. Abstractly, vectors in Einstein notation are elements of a V, with indices serving as placeholders for components in a chosen basis without specifying the basis itself, enabling coordinate-independent expressions; free indices in such equations maintain consistency across terms. This framework underscores the notation's power in generalizing linear algebra to curved spaces or non-orthogonal bases.

Matrices as Tensors

In Einstein notation, matrices are represented as second-order mixed tensors of type (1,1), denoted as A^i_j, where the contravariant index i corresponds to the row and the covariant index j to the column, encapsulating a between vector spaces. This representation aligns with the tensor's role in vectors while respecting index placement conventions. The component A^i_j specifies how the basis vector \mathbf{e}_j is transformed, yielding A \mathbf{e}_j = A^i_j \mathbf{e}_i, where the summation over the repeated index i is implied by the Einstein convention. This access highlights the matrix's action on basis elements, extending the vector notation where single indices denote components in a chosen basis. Matrices in this framework are typically of mixed valence, combining one contravariant and one covariant index; to obtain a fully covariant form A_{ij}, the contravariant index is lowered using the metric tensor via A_{ij} = g_{ik} A^k_j, with summation over k. This operation preserves the tensor's geometric meaning while adapting it to all-covariant representations, such as those used in inner products. Under a change of basis, the components transform according to the rule A'^i_j = \frac{\partial x'^i}{\partial x^k} A^k_l \frac{\partial x^l}{\partial x'^j}, ensuring the tensor's invariance as a multilinear object. This mixed transformation law reflects the differing behaviors of contravariant and covariant indices. In orthonormal bases, such as Cartesian coordinates, the distinction between upper and lower indices diminishes due to the Kronecker delta metric (\delta^i_j), allowing indices to be written equivalently as all upper or all lower without altering the components. This simplification facilitates computations in Euclidean spaces. More generally, matrices relate to multilinear maps; a (1,1)-tensor corresponds to a , but by adjusting indices—such as lowering both to form a (0,2)-tensor—it represents a , taking two vectors and yielding a scalar independent of basis choice. This perspective underscores the tensorial nature of matrix operations in varied index configurations.

Key Operations and Manipulations

Contractions and Inner Products

In Einstein notation, a is a summation operation over a pair of repeated indices, one contravariant (upper) and one covariant (lower), which reduces the of a tensor by two. For instance, given a mixed tensor T^i_{j k}, the T^i_{i k} yields \sum_i T^i_{i k}, resulting in a covariant (rank 0,1). This operation is basis-independent and preserves the tensorial nature of the expression under coordinate transformations. The inner product of two vectors provides a fundamental example of contraction, producing a scalar invariant. In Euclidean space, the inner product of vectors \mathbf{u} and \mathbf{v} is denoted \mathbf{u} \cdot \mathbf{v} = u_i v^i = \sum_i u_i v^i, where the indices align for summation. In more general settings, such as Riemannian manifolds, the metric tensor g_{ij} facilitates the contraction: \mathbf{u} \cdot \mathbf{v} = g_{ij} u^i v^j = \sum_{i,j} g_{ij} u^i v^j, ensuring the result is independent of the basis choice. This extends to higher-rank tensors by contracting compatible index pairs, reducing the overall order while maintaining invariance. For second-rank tensors, the trace exemplifies a full contraction over both indices, yielding a scalar: \operatorname{tr}(A) = A^i_i = \sum_i A^i_i, which sums the diagonal elements in a chosen basis and equals the sum of eigenvalues. More generally, a double contraction on a rank-4 tensor, such as T^{ij}_{kl}, can be written as T^{ij}_{ij} = \sum_{i,j} T^{ij}_{ij}, producing a scalar invariant. Properties of contractions include the commutativity of dummy indices—reordering summed indices does not alter the result—and relabeling invariance, where renaming dummies (e.g., i to m) leaves the expression unchanged, as the summation range is implicit. In physics, contractions appear prominently in general relativity, where the stress-energy tensor T_{\mu\nu} encodes energy-momentum distribution. The energy density \rho observed by an inertial frame with four-velocity u^\mu is obtained via contraction: \rho = T_{\mu\nu} u^\mu u^\nu = \sum_{\mu,\nu} T_{\mu\nu} u^\mu u^\nu, normalizing to the rest-frame energy for a . This scalar measures local energy content and couples to the .

Outer Products and Kronecker Delta

In Einstein notation, the outer product, also known as the tensor product, provides a means to construct higher-rank tensors from lower-rank ones without invoking summation over shared indices. For two vectors \mathbf{u} (with contravariant components u^i) and \mathbf{v} (with covariant components v_j), the outer product is denoted \mathbf{u} \otimes \mathbf{v} and yields a (1,1)-tensor with components u^i v_j, where the indices remain free and no summation is implied. This operation increases the total rank of the resulting tensor by the sum of the ranks of the input tensors, preserving their individual index types. More generally, the outer product of two tensors, such as a contravariant vector a^i and another contravariant vector b^j, produces a rank-2 contravariant tensor T^{ij} = a^i b^j. For instance, the outer product of a rank-1 tensor and a rank-2 tensor results in a rank-3 tensor, with components formed by simple multiplication of corresponding index expressions. This construction is fundamental in tensor algebra, as it allows the systematic building of multilinear maps from basic vectorial quantities, with the free indices dictating the tensor's transformation properties under coordinate changes. The , denoted \delta^i_j, is a mixed (1,1)-tensor defined such that \delta^i_j = 1 if i = j and \delta^i_j = 0 otherwise, serving as the in tensor . In Einstein notation, it functions as an for manipulations, exemplified by the \delta^i_j v^j = v^i, where the over the repeated j effectively selects the i-th component of the \mathbf{v}. This property underscores its role in simplifying expressions and representing the identity transformation in linear algebra contexts. Key properties of the Kronecker delta include its orthogonality and composition rules, such as \sum_k \delta^i_k \delta^k_j = \delta^i_j, which follows from its nature and confirms that repeated application yields the original delta. This summation property highlights its utility in verifying tensor identities and in applications requiring orthogonal projections. For example, the Kronecker delta appears in the construction of projection operators, such as the orthogonal projector onto a spanned by basis vectors, where components involve \delta^i_j to isolate directions. Similarly, in change-of-basis matrices, it facilitates the representation of the map between coordinate systems, ensuring invariance under transformations.

Index Raising and Lowering

In Einstein notation, the metric tensor plays a central role in converting between contravariant and covariant representations of tensors by raising or lowering indices. The covariant metric tensor g_{ij}, which is symmetric and provides the inner product structure on the space, lowers a contravariant index on a vector v^j to yield its covariant form via the contraction v_i = g_{ij} v^j, where summation over the repeated index j is implied. Conversely, the inverse metric tensor g^{ij}, satisfying g^{ik} g_{kj} = \delta^i_j, raises a covariant index to produce the contravariant form v^i = g^{ij} v_j. These operations preserve the tensorial nature and are essential for maintaining consistency in index manipulations across different coordinate systems. In Euclidean space, where the geometry is flat and positive definite, the metric tensor simplifies to the Kronecker delta g_{ij} = \delta_{ij}, making the raising and lowering operations trivial identities: v_i = v^i and vice versa, as no sign changes or scaling occur. This equivalence highlights that, in such spaces, the distinction between covariant and contravariant components is often nominal unless higher-order tensors are involved. For mixed tensors, such as a (1,1) tensor A^i_j, raising or lowering proceeds component-wise; for instance, lowering the upper index gives A_{k j} = g_{k i} A^i_j, or more generally A^i_j = g^{ik} A_{k j} to adjust the lower index, ensuring that contractions remain invariant under these transformations. In the context of special relativity, the Minkowski metric \eta_{\mu\nu} = \operatorname{diag}(-1,1,1,1) introduces a pseudo-Euclidean signature, where raising and lowering indices can introduce sign flips depending on whether the index is timelike or spacelike, affecting the physical interpretation of quantities like four-vectors. For the operations to be well-defined and invertible, the metric tensor must be non-degenerate, meaning its determinant is non-zero, which guarantees the existence of the inverse and allows reversible index conversions without loss of information. A practical example arises in differential geometry, where the gradient of a scalar function f is naturally a covariant vector \nabla_i f = \partial_i f, representing the components in the dual basis; raising the index yields the contravariant gradient \nabla^i f = g^{ij} \nabla_j f, which aligns with the directional derivative along coordinate lines.

Advanced Applications

In Tensor Calculus

In tensor calculus, Einstein notation provides a compact way to express derivatives on manifolds. The partial derivative operator is denoted as \partial_i = \frac{\partial}{\partial x^i}, transforming as a covariant vector under coordinate changes. For a scalar function f, higher-order partial derivatives appear as \partial_i \partial_j f, where the indices indicate the components of the Hessian tensor, with summation implied only over repeated indices in more complex expressions. To account for curvature, the covariant derivative replaces the partial derivative while preserving tensor transformation properties. For a (1,1) tensor T^i_j, it is defined as \nabla_k T^i_j = \partial_k T^i_j + \Gamma^i_{k l} T^l_j - \Gamma^l_{k j} T^i_l, where \Gamma^i_{kl} are the Christoffel symbols, which are constructed from the metric tensor via index raising and lowering. This formula generalizes to tensors of arbitrary rank by adding connection terms with positive signs for contravariant indices and negative signs for covariant ones. The Riemann curvature tensor captures the intrinsic geometry of the manifold and is expressed using Einstein notation as R^\rho_{\ \sigma \mu \nu} = \partial_\mu \Gamma^\rho_{\nu \sigma} - \partial_\nu \Gamma^\rho_{\mu \sigma} + \Gamma^\rho_{\mu \lambda} \Gamma^\lambda_{\nu \sigma} - \Gamma^\rho_{\nu \lambda} \Gamma^\lambda_{\mu \sigma}. This (1,3) tensor measures the failure of parallel transport around closed loops. Contracting the first and third indices yields the Ricci tensor, R_{\mu \nu} = R^\lambda_{\ \mu \lambda \nu}, a symmetric (0,2) tensor describing average curvature. Further contraction with the inverse metric produces the Ricci scalar, R = g^{\mu \nu} R_{\mu \nu}, a coordinate-invariant measure of overall curvature. Einstein notation streamlines these curvature expressions in differential geometry, enabling efficient manipulation of multi-index objects without explicit summation symbols. It particularly compacts derivations in general relativity, where such tensors underpin geometric computations. A refinement known as abstract index notation, introduced by Roger Penrose, uses indices to specify tensor type (e.g., T^a_b for a (1,1) tensor) without implying a coordinate basis, promoting a more abstract, coordinate-free perspective in tensor calculus.

In Physics Contexts

Einstein notation finds extensive application in physics, particularly in theories involving spacetime and multi-dimensional fields, where it facilitates the compact expression of tensor equations and ensures covariance under transformations. In special relativity, four-vectors such as the position four-vector x^\mu = (ct, \mathbf{x}) transform under Lorentz boosts via x'^\mu = \Lambda^\mu_\nu x^\nu, where repeated indices imply summation over \mu, \nu = 0, 1, 2, 3. The invariant spacetime interval is given by ds^2 = \eta_{\mu\nu} dx^\mu dx^\nu, with the Minkowski metric \eta_{\mu\nu} = \operatorname{diag}(-1, 1, 1, 1), highlighting how index notation preserves Lorentz invariance without explicit sums. In general relativity, Einstein notation is essential for formulating the field equations that relate spacetime curvature to matter distribution. The Einstein field equations are G_{\mu\nu} = \frac{8\pi G}{c^4} T_{\mu\nu}, where the Einstein tensor is G_{\mu\nu} = R_{\mu\nu} - \frac{1}{2} g_{\mu\nu} R, with R_{\mu\nu} the Ricci tensor, R the Ricci scalar, and g_{\mu\nu} the metric tensor; this index form encapsulates the non-linear coupling of geometry and energy-momentum. Electromagnetism benefits from Einstein notation through the antisymmetric Faraday tensor F^{\mu\nu} = \partial^\mu A^\nu - \partial^\nu A^\mu, derived from the four-potential A^\mu = (\phi/c, \mathbf{A}), which unifies electric and magnetic fields in a Lorentz-covariant manner. One of Maxwell's equations in this framework is \partial_\mu F^{\mu\nu} = \mu_0 J^\nu, where J^\nu is the four-current, demonstrating how contraction over indices yields the source terms succinctly. In , Einstein notation aligns with Dirac's bra-ket formalism by expressing the inner product as \langle \psi | \phi \rangle = \psi^*_i \phi^i, a summation over basis indices that computes the overlap between states. Density matrices, representing mixed states, are likewise written in component form as \rho_{ij} = \sum_k p_k \psi^{(k)*}_i \psi^{(k)}_j, where p_k are probabilities, enabling efficient handling of ensembles via matrix indices. Beyond these core areas, Einstein notation appears in through the \sigma^{ij}, which describes and appears in the Navier-Stokes equations as \partial_j \sigma^{ij} + \rho f^i = \rho \frac{D v^i}{Dt}, capturing viscous and forces in index-contracted . In , routinely employ indices, such as the Dirac Lagrangian \mathcal{L} = \bar{\psi} (i \gamma^\mu \partial_\mu - m) \psi, where over Lorentz indices \mu ensures relativistic invariance in dynamics. The advantages of Einstein notation in physics stem from its ability to handle spacetime indices seamlessly, making multi-dimensional summations implicit and reducing algebraic errors in covariant expressions across theories from to field theory. This convention streamlines the verification of physical laws under coordinate transformations, as seen in the unified treatment of vectors and tensors.

References

  1. [1]
    Einstein Summation -- from Wolfram MathWorld
    Einstein summation is a notational convention for simplifying expressions including summations of vectors, matrices, and general tensors.
  2. [2]
    [PDF] The Einstein Summation Notation
    The Einstein summation notation is an algebraic short-hand for expressing multicomponent Carte- sian quantities, manipulating them, simplifying.Missing: history | Show results with:history
  3. [3]
    [PDF] THE FOUNDATION OF THE GENERAL THEORY OF RELATIVITY
    he “HE special theory of relativity is based on the following postulate, which is also satisfied by the mechanics of Galileo and Newton.
  4. [4]
    Einsteins Summation Convention - an overview | ScienceDirect Topics
    The Einstein summation convention is defined as a notational method in tensor mathematics where an index that occurs exactly twice in a tensor expression is ...
  5. [5]
  6. [6]
  7. [7]
    [PDF] Roger Penrose: Applications of Negative Dimensional Tensors
    I take as my model, the conventional tensor index notation with Einstein's summation convention, which has become so familiar in physics and in what is now ...
  8. [8]
    [PDF] The Einstein Summation Notation - Dr. Brian Sullivan
    The Einstein summation notation is an algebraic short-hand which allows multidimensional Carte- sian quantities to be expressed, manipulated, and.<|control11|><|separator|>
  9. [9]
    [PDF] Gravitation - physicsgg
    This is a textbook on gravitation physics (Einstein's "general relativity" or "geo- metrodynamics"). It supplies two tracks through the subject.
  10. [10]
    [PDF] Einstein summation convention and δ-functions
    Einstein summation convention is a convenient notation when manipulating expressions involving vectors, matrices, or tensors in general. (A tensor is a.Missing: history | Show results with:history
  11. [11]
    [PDF] Applications of Negative Dimensional Tensors
    I take as my model, the conventional tensor index notation with Einstein's summation convention, which has become so familiar in physics and in what is now ...
  12. [12]
    [PDF] Primer on Index Notation - DSpace@MIT
    In his presentation of relativity theory, Einstein introduced an index-based notation that has become widely used in physics. This notation is almost ...<|separator|>
  13. [13]
    [PDF] The Einstein convention, indices and networks - Bard Faculty
    Finally we distinguish between free and dummy indices. A free index is one that is not summed over and consequently can take any of the values in its allowed ...
  14. [14]
    [PDF] Einstein's Field Equations
    index used, ν, is a Greek letter. The Greek index is a variable which ... All indices run over the 4 spacetime coordinates (0 = time, 1, 2, 3 = space) ...
  15. [15]
    [PDF] The displacement 4-vector - MIT
    5.8 An aside: Upstairs, downstairs; contravariant, covariant. The use of index notation abounds for representing vectorial quantities (and, more generally,.
  16. [16]
    [PDF] 1. Vectors, contravariant and covariant
    One is called the contravariant vector or just the vector, and the other one is called the covariant vector or dual vector or one-vector.
  17. [17]
    [PDF] Tensors - Purdue Engineering
    – Using Einstein notation, since. – Then. ∇𝐴. 𝑓 = 𝐵 ∇𝜎 𝐴𝑦 + 𝑏 ∇𝐴. 𝐴𝑦. This is a tensor! ∇𝐴. 𝑓. 𝑖1. 𝑗1. ,𝑗2. = 𝐵. 𝑖1. 𝑖2. ∇𝜎. 𝑖2. 𝑖3. ∇𝐴. 𝐴𝑦.Missing: placement | Show results with:placement
  18. [18]
    [PDF] Chapter 10 Vectors and Tensors
    If the eµ constitute an orthonormal basis, where gµν = δµν, then the two sets of components (covariant and contravariant) are numerically coincident.Missing: placement | Show results with:placement
  19. [19]
    [PDF] Notes on Index Notation 1 Einstein Summation Convention
    Index notation is a short-hand method of writing entire systems of equations, based on recognizing consistent patterns in the algebra.
  20. [20]
    [PDF] Introduction to Tensor Calculus for General Relativity - MIT
    The Einstein summation convention is assumed: repeated upper and lower indices are to be summed over their ranges,. e.g., AµBµ ≡ A0B0 + A1B1 + A2B2 + A3B3. Four ...
  21. [21]
    [PDF] Tensor Calculus - HAL
    Nov 3, 2024 · Consequently, any tensor in a Riemannian space with well-defined metric can be cast into covariant or contravariant or mixed forms.10. • In the ...
  22. [22]
    [PDF] tensor products ii - keith conrad
    Einstein summation convention. For example, the most basic contraction V ⊗ V ∨ → K, where Pi,j Ti j ei ⊗ ej 7→ Pi Ti i , is written in this convention as. Ti.
  23. [23]
    None
    ### Summary of Einstein Summation Convention, Contractions, and Inner Products from https://www.utc.edu/document/80261
  24. [24]
    [PDF] Introduction to Tensors
    Einstein's summation convention: repeated upper and lower indices ... Inner product: contraction in conjunction with outer product. e.g.: Again, result ...
  25. [25]
    [PDF] CS 598 EVS: Tensor Computations - RELATE
    tensors, defining einsum (term stems from Einstein's summation convention) tensor contraction einsum diagram inner product w = P i uivi outer product wij ...
  26. [26]
    [PDF] charge conservation; electromagnetism; stress-energy tensor
    Oct 1, 2019 · Tαβ = ρrf uαuβ + Prf (gαβ + uαuβ) . It is important to remember that ρrf and Prf are the fluid's energy density and pressure in its restframe.Missing: Einstein notation
  27. [27]
  28. [28]
    Tensor Notation (Basics) - Continuum Mechanics
    Tensor notation introduces one simple operational rule. It is to automatically sum any index appearing twice from 1 to 3.
  29. [29]
    [PDF] Lecture V: Tensor algebra in flat spacetime
    Raising and lowering indices. Note that because indices can be raised or lowered (see Lecture IV), as long as one has a metric tensor (ηµν in special ...
  30. [30]
    [PDF] An Introduction to Vectors and Tensors from a Computational ... - UTC
    In general, the i and j indices can be assigned separate ranges, for example to represent a 3 5. × matrix. However, all indices are assumed to have the same. 3.Missing: k_j = i_j
  31. [31]
    [PDF] Part 3 General Relativity - DAMTP
    (R4,η) is called Minkowski spacetime. A coordinate chart which covers all of R4 and in which the components of the metric are ηµν ≡ diag(−1,1,1,1).
  32. [32]
    [PDF] Resumé of General Relativity & Gravitation Part I
    j i as a (1,1) tensor. 1.3.3 Raising and lowering indices and contraction. Given any contravariant vector Ai it is possible to define, via the metric tensor ...<|control11|><|separator|>
  33. [33]
    Lecture Notes on General Relativity - S. Carroll
    The Riemann tensor measures that part of the commutator of covariant derivatives which is proportional to the vector field, while the torsion tensor measures ...<|control11|><|separator|>
  34. [34]
    [PDF] Lorentz Transformations in Special Relativity
    Notes 46: Lorentz Transformations where if Xµ is a 4-vector then δ4(X) = δ(X0)δ(X1)δ(X2)δ(X3). (100). Hint: Use the τ-integration to eliminate the delta ...
  35. [35]
    [PDF] 7.1 Transforming energy and momentum between reference frames
    Recall that we found ∆s2 = −c2∆t2 + ∆x2 + ∆y2 + ∆z2 is a Lorentz invariant: all Lorentz frames agree on the value of ∆s2 between two events.
  36. [36]
    [PDF] 5. Electromagnetism and Relativity - DAMTP
    The Maxwell equations are not invariant under Lorentz transformations. This is because there is the dangling ⌫ index on both sides. However, because the ...
  37. [37]
    [PDF] 5. THE ELECTROMAGNETIC FIELD TENSOR
    The anti-symmetric tensor Fµν is called the electromagnetic field tensor; its components will be detailed shortly. • Eq. (7) is the covariant form of the ...
  38. [38]
    [PDF] Linear Algebra In Dirac Notation - CMU Quantum Theory Group
    It is convenient to employ the Dirac symbol |ψi, known as a “ket”, to denote a quantum state without referring to the particular function used to represent it.
  39. [39]
    [PDF] Density Matrices - CMU Quantum Theory Group
    Density matrices are employed in quantum mechanics to give a partial description of a quantum system, one from which certain details have been omitted.
  40. [40]
    [PDF] 3 (More on) The Stress Tensor and the Navier-Stokes Equations
    ... stress tensor (think back to Einstein notation!) and the RHS is the definition of pressure in terms of the elements of the stress tensor for a moving fluid.
  41. [41]
    [PDF] Quantum Field Theory - DAMTP
    tact with general relativity: such an object sits on the right-hand side of Einstein's field equations. In fact this observation provides a quick and easy ...
  42. [42]
    [PDF] EINSTEIN SUMMATION NOTATION
    In addition to the advantage of compactness, writing vectors in this way allows us to manipulate vector calculations and prove vector identities in a much more ...Missing: origin history<|control11|><|separator|>