Fact-checked by Grok 2 weeks ago

Polarization identity

The polarization identity is a fundamental formula in linear algebra that expresses the inner product of two vectors in an in terms of the squared norms of specific linear combinations of those vectors, enabling the recovery of the bilinear or from the associated induced by the norm. In real inner product spaces, the identity states that for any vectors x and y, \langle x, y \rangle = \frac{1}{4} \left( \|x + y\|^2 - \|x - y\|^2 \right). This form arises from expanding the squared norms using the inner product properties and solving for \langle x, y \rangle. For complex inner product spaces, where the inner product is , the identity extends to \langle x, y \rangle = \frac{1}{4} \left( \|x + y\|^2 - \|x - y\|^2 + i \|x + i y\|^2 - i \|x - i y\|^2 \right), incorporating imaginary units to account for the conjugate linearity in the second argument. Alternatively, in the complex case, it can be written as \langle x, y \rangle = \frac{1}{4} \left( \|x + y\|^2 - \|x - y\|^2 \right) + \frac{i}{4} \left( \|x + i y\|^2 - \|x - i y\|^2 \right). The polarization identity plays a key role in characterizing inner product spaces among normed spaces, as it implies that a norm satisfying the parallelogram law \|x + y\|^2 + \|x - y\|^2 = 2(\|x\|^2 + \|y\|^2) arises from an inner product, which can then be reconstructed via the identity. This connection is essential in and , where it facilitates the study of Hilbert spaces and without direct access to the inner product. Generalizations exist for spaces over quaternions, Clifford algebras, and other structures with involutions, often derived by averaging over suitable subgroups to polarize the .

Polarization Identities

In Real Inner Product Spaces

A real is a V over the field of real numbers \mathbb{R} equipped with an inner product \langle \cdot, \cdot \rangle: V \times V \to \mathbb{R}, which is a satisfying \langle x, x \rangle \geq 0 for all x \in V with equality if and only if x = 0. The inner product induces a \|x\| = \sqrt{\langle x, x \rangle}. While the following discussion assumes finite-dimensional spaces for simplicity, the polarization identity extends to infinite-dimensional Hilbert spaces, which are complete with respect to this norm. The polarization identity in a real inner product space expresses the inner product in terms of the : \langle x, y \rangle = \frac{1}{4} \left( \|x + y\|^2 - \|x - y\|^2 \right) for all x, y \in V. An alternative equivalent form is \langle x, y \rangle = \frac{1}{2} \left( \|x + y\|^2 - \|x\|^2 - \|y\|^2 \right). To derive the first form, expand the norms using the inner product properties: \|x + y\|^2 = \langle x + y, x + y \rangle = \langle x, x \rangle + 2 \langle x, y \rangle + \langle y, y \rangle = \|x\|^2 + 2 \langle x, y \rangle + \|y\|^2, \|x - y\|^2 = \langle x - y, x - y \rangle = \langle x, x \rangle - 2 \langle x, y \rangle + \langle y, y \rangle = \|x\|^2 - 2 \langle x, y \rangle + \|y\|^2, where \langle y, x \rangle = \langle x, y \rangle is used. Subtracting these equations yields \|x + y\|^2 - \|x - y\|^2 = 4 \langle x, y \rangle, so dividing by 4 gives the identity. The alternative form follows directly from rearranging the expansion of \|x + y\|^2. As an example, consider \mathbb{R}^2 with the standard Euclidean inner product \langle (a,b), (c,d) \rangle = ac + bd. For x = (1,0) and y = (0,1), compute \|x + y\|^2 = \|(1,1)\|^2 = 2 and \|x - y\|^2 = \|(1,-1)\|^2 = 2. Then \langle x, y \rangle = \frac{1}{4}(2 - 2) = 0, verifying that the standard basis vectors are orthogonal.

In Complex Inner Product Spaces

In complex inner product spaces, the inner product \langle \cdot, \cdot \rangle is sesquilinear, conjugate-linear in the first argument and linear in the second argument. For vectors x, y, z in the space and scalars \alpha, \beta \in \mathbb{C}, this means \langle \alpha x + \beta y, z \rangle = \overline{\alpha} \langle x, z \rangle + \overline{\beta} \langle y, z \rangle and \langle x, \alpha y + \beta z \rangle = \alpha \langle x, y \rangle + \beta \langle x, z \rangle, along with conjugate symmetry \langle x, y \rangle = \overline{\langle y, x \rangle} and positive-definiteness \langle x, x \rangle \geq 0 with equality if and only if x = 0. The polarization identity expresses the inner product in terms of the induced norm \| \cdot \|, where \|x\|^2 = \langle x, x \rangle. For this convention, it takes the form \langle x, y \rangle = \frac{1}{4} \left( \|x + y\|^2 - \|x - y\|^2 - i \|x + i y\|^2 + i \|x - i y\|^2 \right). A compact variant is \langle x, y \rangle = \frac{1}{4} \sum_{k=0}^3 i^k \| x + (-i)^k y \|^2. These formulas recover the inner product from the norm, extending the real case by accounting for the imaginary component through rotations by i. To derive the identity, expand each norm squared using the sesquilinear properties. Begin with the real part: \|x + y\|^2 = \langle x + y, x + y \rangle = \|x\|^2 + \|y\|^2 + \langle x, y \rangle + \langle y, x \rangle = \|x\|^2 + \|y\|^2 + 2 \operatorname{Re} \langle x, y \rangle, \|x - y\|^2 = \|x\|^2 + \|y\|^2 - 2 \operatorname{Re} \langle x, y \rangle. Subtracting yields \|x + y\|^2 - \|x - y\|^2 = 4 \operatorname{Re} \langle x, y \rangle. For the imaginary part, compute \|x + i y\|^2 = \langle x + i y, x + i y \rangle = \|x\|^2 + i \langle x, y \rangle - i \langle y, x \rangle + \|y\|^2 = \|x\|^2 + \|y\|^2 + i (\langle x, y \rangle - \overline{\langle x, y \rangle}) = \|x\|^2 + \|y\|^2 - 2 \operatorname{Im} \langle x, y \rangle, where the coefficient of \langle x, y \rangle follows from \langle i y, x \rangle = \overline{i} \langle y, x \rangle = -i \overline{\langle x, y \rangle} and linearity in the second argument for \langle x, i y \rangle = i \langle x, y \rangle. Similarly, \|x - i y\|^2 = \|x\|^2 + \|y\|^2 + 2 \operatorname{Im} \langle x, y \rangle. Subtracting gives \|x - i y\|^2 - \|x + i y\|^2 = 4 \operatorname{Im} \langle x, y \rangle. Thus, \frac{1}{4} (-i \|x + i y\|^2 + i \|x - i y\|^2) = i \operatorname{Im} \langle x, y \rangle. Combining this imaginary contribution with the real part \frac{1}{4} (\|x + y\|^2 - \|x - y\|^2) = \operatorname{Re} \langle x, y \rangle isolates \langle x, y \rangle. The compact sum arises from averaging over the fourth roots of unity, weighted appropriately. As an example, consider \mathbb{C} with the standard inner product \langle z, w \rangle = \overline{z} w, which is conjugate-linear in the first and linear in the second. For x = 1 and y = i, direct gives \langle 1, i \rangle = \overline{1} \cdot i = i. Applying the polarization identity: \|1 + i\|^2 = |1 + i|^2 = 2, \quad \|1 - i\|^2 = 2, \quad \|1 + i \cdot i\|^2 = |1 - 1|^2 = 0, \quad \|1 - i \cdot i\|^2 = |1 + 1|^2 = 4. Thus, \langle 1, i \rangle = \frac{1}{4} (2 - 2 - i \cdot 0 + i \cdot 4) = \frac{1}{4} (4i) = i, verifying the .

Reconstructing the Inner Product

The polarization identity provides an explicit method to recover the inner product from the associated in inner product spaces. Given the \| \cdot \| induced by the inner product \langle \cdot, \cdot \rangle, where \|x\|^2 = \langle x, x \rangle, the identity expresses \langle x, y \rangle solely in terms of norms of linear combinations of x and y. In real inner product spaces, the reconstruction is given by \langle x, y \rangle = \frac{1}{4} \left( \|x + y\|^2 - \|x - y\|^2 \right). This formula arises from expanding the squared norms using the properties of the inner product and solving for the cross term. For complex inner product spaces, where the inner product is sesquilinear, the full reconstruction requires accounting for the imaginary part: \langle x, y \rangle = \frac{1}{4} \left( \|x + y\|^2 - \|x - y\|^2 \right) + \frac{i}{4} \left( \|x + i y\|^2 - \|x - i y\|^2 \right). Equivalently, \langle x, y \rangle = \frac{1}{4} \left( \|x + y\|^2 - \|x - y\|^2 + i \|x + i y\|^2 - i \|x - i y\|^2 \right). This reconstruction is possible because the norm derives from a quadratic form associated with the inner product. In the broader context of normed spaces, the polarization identity is used alongside the to define an inner product when the norm satisfies \|x + y\|^2 + \|x - y\|^2 = 2(\|x\|^2 + \|y\|^2) for all x, y, confirming the space is isometric to an .

Applications

Parallelogram Law and Normed Spaces

The parallelogram law states that in a normed vector space (V, \|\cdot\|), for all vectors u, v \in V, \|u + v\|^2 + \|u - v\|^2 = 2\|u\|^2 + 2\|v\|^2. Geometrically, this identity reflects the property of parallelograms in Euclidean geometry, where the sum of the squares of the diagonals equals twice the sum of the squares of the adjacent sides; in vector terms, it equates the squared lengths of the diagonals of the parallelogram formed by u and v to twice the sum of the squared lengths of u and v themselves. In inner product spaces, the parallelogram law holds as a direct consequence of the norm definition \|x\|^2 = \langle x, x \rangle. Expanding both sides using the inner product yields \|u + v\|^2 + \|u - v\|^2 = \langle u + v, u + v \rangle + \langle u - v, u - v \rangle = 2\langle u, u \rangle + 2\langle v, v \rangle = 2\|u\|^2 + 2\|v\|^2, confirming the equality. Conversely, if a norm on a real vector space satisfies the parallelogram law, it arises from an inner product, recoverable via the polarization identity: \langle u, v \rangle = \frac{1}{4} \left( \|u + v\|^2 - \|u - v\|^2 \right). This expression is bilinear, symmetric, and positive definite, as verified by substituting the law and checking properties like \langle u, u \rangle = \|u\|^2 > 0 for u \neq 0. For complex spaces, a similar but adjusted polarization identity (involving additional terms with i) reconstructs the sesquilinear inner product. The full equivalence—that a norm derives from an inner product if and only if the parallelogram law holds—was established by Jordan and von Neumann. In the context of Banach spaces (complete normed spaces), those satisfying the parallelogram law are precisely the Hilbert spaces, where the inner product induces a complete norm. This characterization distinguishes Hilbert spaces from general Banach spaces like \ell^1, which fail the law. For example, in Euclidean space \mathbb{R}^2 with the \ell^2 norm, taking standard basis vectors e_1 = (1,0) and e_2 = (0,1), we have \|e_1\| = \|e_2\| = 1, \|e_1 + e_2\| = \|(1,1)\| = \sqrt{2}, and \|e_1 - e_2\| = \|(1,-1)\| = \sqrt{2}. The left side is (\sqrt{2})^2 + (\sqrt{2})^2 = 4, and the right side is $2(1)^2 + 2(1)^2 = 4, so the law holds. In contrast, with the \ell^1 (taxicab) norm, \|e_1\|_1 = \|e_2\|_1 = 1, but \|e_1 + e_2\|_1 = \|(1,1)\|_1 = 2 and \|e_1 - e_2\|_1 = 2. The left side is $4 + 4 = 8, while the right side is $2 + 2 = 4, violating the law. In inner product spaces satisfying the , (\langle x, y \rangle = 0) implies the : \|x + y\|^2 = \|x\|^2 + \|y\|^2, as the polarization identity sets the cross term to zero, reducing the expansion directly.

Geometric Interpretations

The polarization identity bridges norms and inner products in , enabling the derivation of angular relations from length measurements alone. In real inner product spaces, the identity expresses the inner product as \langle \mathbf{u}, \mathbf{v} \rangle = \frac{1}{4} \left( \|\mathbf{u} + \mathbf{v}\|^2 - \|\mathbf{u} - \mathbf{v}\|^2 \right), which geometrically interprets the alignment of vectors through observable distances. A key derivation arises from expanding the norm of the vector difference, yielding the law of cosines. Specifically, \|\mathbf{u} - \mathbf{v}\|^2 = \|\mathbf{u}\|^2 + \|\mathbf{v}\|^2 - 2 \langle \mathbf{u}, \mathbf{v} \rangle. Substituting the definition \langle \mathbf{u}, \mathbf{v} \rangle = \|\mathbf{u}\| \|\mathbf{v}\| \cos \theta, where \theta is the angle between \mathbf{u} and \mathbf{v}, gives \|\mathbf{u} - \mathbf{v}\|^2 = \|\mathbf{u}\|^2 + \|\mathbf{v}\|^2 - 2 \|\mathbf{u}\| \|\mathbf{v}\| \cos \theta. This links norms directly to the cosine of the included , providing a foundational geometric in spaces. In , the polarization identity facilitates computation without coordinate representations, using solely the norms of sums and differences as proxies for directional information. This approach emphasizes the identity's role in recovering scalar projections and angular measures from metric data, aligning with classical geometric constructions where lengths determine configurations. In \mathbb{R}^3, an illustrative computation of the angle between vectors \mathbf{u} and \mathbf{v} relies only on lengths of sums and differences. Suppose \|\mathbf{u}\| = 4, \|\mathbf{v}\| = 4, \|\mathbf{u} + \mathbf{v}\| = 5, and \|\mathbf{u} - \mathbf{v}\| = 3; then \langle \mathbf{u}, \mathbf{v} \rangle = \frac{1}{4} (25 - 9) = 4, so \cos \theta = \frac{4}{16} = 0.25 and \theta \approx 75.52^\circ. This method highlights the identity's utility in spatial for determination from observations.

Isometries in Inner Product Spaces

In inner product spaces, an is defined as a that preserves the of every , meaning \|T(x)\| = \|x\| for all x in the space. This preservation of norms extends to the inner product through the polarization identity, which reconstructs the inner product from the ; thus, if T is linear and norm-preserving, then \langle T(x), T(y) \rangle = \langle x, y \rangle for all x, y. In real inner product spaces, such isometries are precisely the orthogonal transformations, which satisfy T^T T = I, where T^T is the (transpose in the case), ensuring the inner product is preserved. For example, in \mathbb{R}^2 with the standard inner product, a \begin{pmatrix} \cos \theta & -\sin \theta \\ \sin \theta & \cos \theta \end{pmatrix} is an that preserves norms and thus the inner product via the real polarization identity \langle x, y \rangle = \frac{1}{4} (\|x + y\|^2 - \|x - y\|^2). In complex inner product spaces, linear isometries are unitary operators, satisfying T^* T = I, where T^* is the , and they preserve the sesquilinear inner product using the complex polarization identity \langle x, y \rangle = \frac{1}{4} (\|x + y\|^2 - \|x - y\|^2 + i \|x + i y\|^2 - i \|x - i y\|^2). In Hilbert spaces, linear isometries preserve inner products via the polarization identity and are unitary operators if surjective; more generally, they coincide with unitary operators up to anti-unitary components in the broader .

Generalizations

Jordan–von Neumann Theorems

The theorem provides a characterization of norms induced by inner products in . Specifically, a over the real or complex field is an if and only if its norm satisfies the : for all vectors x, y in the space, \|x + y\|^2 + \|x - y\|^2 = 2\|x\|^2 + 2\|y\|^2. In the setting of Banach spaces, which are complete normed spaces, this condition is equivalent to the space being a , as the induced inner product ensures completeness with respect to the given norm. The proof relies centrally on the to reconstruct the inner product from the norm. Assuming the holds, one defines a using the polarization formula—for the real case, \langle x, y \rangle = \frac{1}{4} \left( \|x + y\|^2 - \|x - y\|^2 \right), with an analogous expression involving additional terms for the complex case; the parallelogram law then guarantees that this form is symmetric (or Hermitian), positive definite, and recovers the original norm via \|x\|^2 = \langle x, x \rangle. The converse direction follows directly from the properties of inner products, which always satisfy the parallelogram law. This result was established by and in their joint 1935 paper, where they developed the for linear spaces. Extensions of the include approximate versions for spaces that nearly satisfy the , quantified by the von Neumann–Jordan constant C_{NJ}(X) = \sup_{x,y \neq 0} \frac{\|x+y\|^2 + \|x-y\|^2}{2(\|x\|^2 + \|y\|^2)}, which equals 1 if and only if X is an ; this constant is particularly useful in uniformly convex Banach spaces to measure deviation from Hilbert structure. Other characterizations link the theorem to uniform convexity: a uniformly convex Banach space is Hilbert if its modulus of convexity matches that of Hilbert spaces, aligning with the parallelogram condition. A classic example is the sequence space \ell^2, where the norm satisfies the parallelogram law and the space is Hilbert with the standard inner product; in contrast, \ell^p for $1 \leq p < \infty with p \neq 2 fails the law—for instance, taking standard basis vectors shows the equality does not hold—and thus is not Hilbert.

Symmetric Bilinear Forms

In the algebraic setting, a quadratic form q on a V over a F of characteristic not equal to 2 is a map q: V \to F satisfying q(\lambda v) = \lambda^2 q(v) for all \lambda \in F and v \in V. Such a quadratic form induces a unique b: V \times V \to F, defined by the polarization identity b(x, y) = \frac{1}{4} \bigl( q(x + y) - q(x - y) \bigr). This bilinear form satisfies q(v) = b(v, v) and b(\lambda x, y) = \lambda b(x, y) = b(x, \lambda y) for all scalars \lambda, with symmetry b(x, y) = b(y, x). The inner product on a real represents a special case where b is positive definite. To verify the polarization identity, expand the right-hand side using the bilinearity of b (which holds once defined this way, with uniqueness ensured in characteristic not 2): q(x + y) = b(x + y, x + y) = b(x, x) + 2b(x, y) + b(y, y), q(x - y) = b(x - y, x - y) = b(x, x) - 2b(x, y) + b(y, y). Subtracting these equations yields q(x + y) - q(x - y) = 4b(x, y), so dividing by 4 recovers b(x, y). This expansion confirms that the formula is consistent and that b is indeed symmetric and bilinear. The assumption of characteristic not 2 is essential, as division by 4 (or equivalently by 2) would fail otherwise, preventing unique recovery of b from q. In characteristic 2, alternative approaches are needed to relate quadratic and bilinear forms. In algebraic applications, the polarization identity facilitates the decomposition and classification of quadratic forms over fields such as the rationals or finite fields, by allowing the extraction of the full symmetric bilinear structure from diagonal evaluations alone. For instance, it enables the study of orthogonality and isotropy in quadratic spaces, where subspaces are analyzed via the associated bilinear form. A example arises in the matrix representation of . Consider q(x) = x^T A x for x \in F^n and A = (a_{ij}), so q(e_i + e_j) = a_{ii} + 2a_{ij} + a_{jj} and q(e_i - e_j) = a_{ii} - 2a_{ij} + a_{jj}, where e_k are vectors. The polarization identity then gives a_{ij} = b(e_i, e_j) = \frac{1}{4} \bigl( q(e_i + e_j) - q(e_i - e_j) \bigr) for i \neq j, directly recovering the off-diagonal entries from the quadratic form evaluations.

Higher-Degree Homogeneous Polynomials

The polarization identity generalizes to higher-degree homogeneous polynomials, providing a method to recover a unique symmetric n-linear form B from a degree-n homogeneous polynomial p on a vector space over a field, where p(x) = B(x, \dots, x). This extension applies in settings beyond the quadratic case, which serves as the n=2 instance for bilinear forms. The process relies on techniques such as finite differences or directional derivatives to extract the multilinear structure from the diagonal evaluations of p. A standard polarization formula uses iterated forward difference operators \Delta_h p(x) = p(x + h) - p(x), applied successively and scaled appropriately. Specifically, the symmetric n-linear form is given by B(x_1, \dots, x_n) = \frac{1}{n!} \operatorname{Tr} \left( \Delta_{x_n} \cdots \Delta_{x_1} p \right), where \operatorname{Tr} denotes evaluation at the origin. An equivalent combinatorial expression, derived via inclusion-exclusion, is n! \, B(x_1, \dots, x_n) = \sum_{k=0}^n (-1)^{n-k} \sum_{|J|=k} p\left( \sum_{i \in J} x_i \right), with the inner sum ranging over all subsets J \subseteq \{1, \dots, n\} of cardinality k. These formulas involve averaging over permutations or subsets to symmetrize the recovery, ensuring B is fully multilinear. The approach requires the underlying field to have characteristic zero or coprime to n, as the presence of n! in the denominator demands its invertibility; otherwise, the mapping may not be bijective. For illustration, consider the cubic case where n=3 and p(x) = \langle x, x, x \rangle for some trilinear form \langle \cdot, \cdot, \cdot \rangle. The polarization yields the full symmetric trilinear form via B(x,y,z) = \frac{1}{6} \left[ p(x+y+z) - p(x+y) - p(y+z) - p(z+x) + p(x) + p(y) + p(z) \right], explicitly reconstructing \langle x, y, z \rangle (up to symmetrization) from the cubic diagonal. This demonstrates how higher-degree polarizations disentangle the multilinear components embedded in the . These generalizations play a key role in , where polarization decomposes homogeneous invariant polynomials into multilinear invariants under group actions, facilitating the study of invariant rings and equivariant maps. In , they aid in analyzing polynomial representations of groups like \mathrm{GL}(V), by relating homogeneous components to their multilinear polarizations and enabling proofs of fundamental theorems on invariant generation.

Extensions to Other Algebraic Structures

Further generalizations of the polarization identity extend to inner product spaces over non-commutative division algebras, such as the (\mathbb{H}), and to more abstract structures like Clifford algebras. In these settings, the inner product is replaced by a Hermitian form compatible with the algebra's (conjugation). The polarization is obtained by averaging the quadratic form over a suitable compact subgroup of the using the , ensuring the recovery of the . For example, over quaternions, the identity involves terms that account for the non-commutativity, similar to the case but with quaternion conjugation. These extensions preserve the connection to the in the generalized theorem for such algebras.

Historical Context

Classical Origins

The classical origins of polarization ideas lie in the 18th-century study of forms within and , where mathematicians sought to decompose and analyze these homogeneous polynomials of two. Joseph-Louis Lagrange's 1759 method for reducing forms to sums of squares marked a pivotal advancement, employing to simplify expressions and compute discriminants, which reveal the form's signature and solvability properties. This approach, applied to binary and ternary forms, facilitated foundational derivations for representing integers and solving Diophantine equations, without naming a specific but establishing techniques to extract cross terms akin to bilinear interactions. Leonhard Euler's contributions in during the mid-18th century provided essential groundwork through his development of homogeneous functions, as detailed in his (1748). Euler explored forms as special cases of these functions, linking them to geometric problems like sums of squares and using algebraic manipulations—such as expansions involving linear combinations—to investigate their properties in . His experimental studies on forms like x^2 + y^2 and connections to emphasized discriminant computations for classification, influencing Lagrange's systematic reductions and highlighting forms' role in broader analytic contexts. Carl Friedrich Gauss elevated these ideas in his Disquisitiones Arithmeticae (1801), offering a rigorous theory of binary quadratic forms in and . Gauss employed derivations that polarized forms by considering linear combinations to define , , and class numbers, enabling geometric interpretations such as reductions and reciprocity laws. These methods, rooted in for analysis, formed the algebraic backbone for later bilinear recoveries, though no explicit identity was articulated. The transition to vector interpretations occurred in the early , as incorporated bilinear forms into analytic frameworks around 1821, and William Rowan Hamilton's quaternions (1843) introduced operations that echoed symmetries in higher dimensions. This evolution connected classical techniques to emerging vector spaces, prefiguring modern inner product structures.

Modern Developments

The polarization identity received its modern formalization in functional analysis during the early 1930s, driven by John von Neumann's investigations in quantum mechanics and operator theory. In his 1932 monograph Mathematical Foundations of Quantum Mechanics, von Neumann introduced abstract Hilbert spaces as the rigorous framework for quantum states, relying on inner products to define observables and employing norm-based derivations that implicitly invoked polarization-like relations. Extending this from 1932 to 1937, von Neumann's papers on unbounded operators and spectral theory in Hilbert spaces explicitly utilized the identity to reconstruct sesquilinear forms from quadratic norms, solidifying its role in infinite-dimensional settings. The term "polarization identity" emerged directly from von Neumann's framework, where it encapsulates the unique recovery of the inner product from the under the , distinguishing inner product spaces from general normed spaces. This naming reflects von Neumann's emphasis on as a bilinear extension of forms, a concept he developed to handle the symmetry and antilinearity in complex s central to quantum . Pascual Jordan's contributions in complemented and intersected with von Neumann's, culminating in their joint 1935 paper establishing the Jordan–von Neumann theorem. This theorem proves that a on a real or complex arises from an inner product it satisfies the , with the polarization identity providing the explicit formula for the inner product: for real spaces, \langle x, y \rangle = \frac{1}{4} (\|x + y\|^2 - \|x - y\|^2), and an analogous complex version involving cyclic sums. Jordan's earlier algebraic explorations into non-associative structures influenced the theorem's proof, highlighting the identity's utility in characterizing metric spaces with bilinear symmetries. In the post-World War II era, the polarization identity proliferated through influential textbooks, embedding it in the core of education. Paul Halmos discussed it extensively in his 1951 text Introduction to Hilbert Space and the Theory of Spectral Multiplicity, using it to illustrate the geometry of and their operator algebras for graduate students. Peter Lax further popularized it in the 1980s via lectures and publications, including precursors to his 2002 book , where he applied the identity to PDEs and approximation theory while stressing its foundational role in normed space theory. Contemporary research continues to refine the , with generalizations maintaining its classical . A 2022 arXiv preprint extends the formula to arbitrary complex inner product spaces, deriving the from norms via a unified expression that accommodates non-standard scalar fields, demonstrating the identity's robustness beyond settings.

References

  1. [1]
    (PDF) The development of inner product spaces and its generalization
    Aug 6, 2025 · An inner product space is a vector space with an additional structure called the inner product. This additional structure associates each ...
  2. [2]
    [2204.12014] Polarization Identities - arXiv
    Apr 26, 2022 · We prove a generalization of the polarization identity of linear algebra expressing the inner product of a complex inner product space in terms of the norm.Missing: original | Show results with:original
  3. [3]
    [PDF] polarization identities - Central Michigan University
    Abstract. We prove a generalization of the polarization identity of linear algebra ex- pressing the inner product of a complex inner product space in terms ...
  4. [4]
    [PDF] MATH 423 Linear Algebra II Lecture 28: Inner product spaces.
    Theorem 2 (Polarization Identity) Suppose V is an inner product space with an inner product (·,·) and the induced norm ‖·‖. (i) If V is a real vector space, ...
  5. [5]
    [PDF] Functional Analysis MAT 602 - Uttarakhand Open University
    Kreyszig, (1989), Introductory Functional Analysis with applications, John Wiley and Sons. ... Theorem 8.4.5(Polarization Identity): Let 𝑋 be an inner product ...
  6. [6]
    [PDF] Real Inner Product Spaces and Orthogonal Transformations Math ...
    In this case we use the notation x • y to denote the inner product. Theorem 4 (The polarization identity) Let x,y be elements of an inner product sapce V. Then.
  7. [7]
    [PDF] 7 Inner Products
    In quantum mechanics and mathematical physics, one traditionally defines linearity in the second argument. • Conjugate-symmetry guarantees that, for all vectors ...
  8. [8]
    [PDF] Hilbert Spaces Definition. A complex inner product space (or pre ...
    Definition. A complex inner product space (or pre-Hilbert space) is a complex vector space X together with an inner product: a function from X x X into C ...
  9. [9]
    [PDF] arXiv:2204.12014v2 [math.RA] 6 Feb 2023
    Feb 6, 2023 · Abstract. We prove a generalization of the polarization identity of linear algebra ex- pressing the inner product of a complex inner product ...Missing: ² - ² + ²)<|separator|>
  10. [10]
    [PDF] Inner Product Spaces - Linear Algebra Done Right
    Prove that a norm satisfying the parallelogram equality comes from an inner product (in other words, show that if k k is a norm on U satisfying the.
  11. [11]
    [PDF] on inner products in linear, metric spaces
    JORDAN AND J. V. NEUMANN all complex numbers a the quantities a·f, ƒ + g e L and the real number || ƒ || are defined, with the following properties: 1. a.f ...
  12. [12]
  13. [13]
    [PDF] Notes on Linear Algebra and Functional Analysis
    Jul 19, 2024 · 2.3.3 The Parallelogram Law and the Polarization Identity . . . . . . . . . . ... Figure 4.10: Geometric interpretation of the adjoint A†.
  14. [14]
    [PDF] Hilbert Spaces I: Basic Properties - KSU Math
    These two identities are referred to as the Law of Cosine. 3 ... The fact that q satisfies the Parallelogram Law (15), as well as the Polarization Identity.
  15. [15]
    [PDF] 2 Coordinate systems - CSCI
    Figure 3.4 The Law of Cosines. 3.9. The angle between two vectors: The Law of Cosines ... Figure 4.5 The polarization decomposition. 4.5. The ambiguity of ...
  16. [16]
    [PDF] 10.3 Linear Isometries (Also Called Unitary Transformations)
    which shows that f preserves the Hermitian inner product, as desired. ... For this we used the polarization identity. 2u · v = ‖u‖2. + ‖v‖2 − ‖u − v‖2. In the ...
  17. [17]
    [PDF] lecture 24: orthogonality and isometries
    For an inner product space, an isometry also preserves the inner product: <v, w> = <Lv, L w>. This is because of the polarization identities which relate the ...
  18. [18]
    [PDF] ORTHOGONAL MATRICES Math 21b, O. Knill
    Orthogonal transformations preserve the length of vectors as well as the angles between them.Missing: polarization | Show results with:polarization
  19. [19]
    Isometries of real Hilbert C⁎-modules - ScienceDirect
    Indeed, complex linear isometries preserve inner products due to the polarization identity (1.1). Conversely, suppose h preserves inner products. Then 〈 h ...
  20. [20]
    [PDF] 5. Hilbert spaces Definition 5.1. Let H be a (complex) vector space. A ...
    Hint: Polarization! Such maps are called unitary; they preserve the complete Hilbert space structure. In other words, we can now say that Theorem 5.14 shows ...
  21. [21]
    [PDF] Chapter 1 Hilbert space and linear operators
    Clearly, a unitary operator is an instance of an isometry. For isometries the following properties can easily be obtained. Proposition 1.3.7. a) Let V ∈ B(H) be ...
  22. [22]
    [PDF] An Approximate Version of the Jordan von Neumann Theorem for ...
    Sep 4, 2013 · Abstract. It is known that any normed vector space which satisfies the paral- lelogram law is actually an inner product space.
  23. [23]
    None
    ### Summary of Content from the PDF
  24. [24]
    [PDF] Lecture 15 - UBC Math
    Prop: LP obeys the paralleogram law iff p = 2. Cor: LP is a Hibert space iff p = 2. "If:" we already know that l² is an inner product space and thus satisfies ...
  25. [25]
    [PDF] BILINEAR FORMS The geometry of Rn is controlled algebraically by ...
    For instance, it shows us that a symmetric bilinear form B is identically 0 if and only if B(v, v) = 0 for all v (not just B(ei,ei) = 0 on a basis; see Example ...
  26. [26]
    [PDF] Supplementary Notes on Linear Algebra - UC Berkeley math
    May 3, 2015 · A nondegenerate positively defined symmetric bilinear pairing on V is called a scalar product. ... known as the Polarization Formula. If a norm ...
  27. [27]
    [PDF] Symmetric bilinear forms
    SYMMETRIC BILINEAR FORMS. The polarization identity (1) shows that a subspace F ⊂ V is isotropic if and only if all of its vectors are isotropic. If F ⊂ V ...
  28. [28]
    [1309.1275] A polarization identity for multilinear maps - arXiv
    Sep 5, 2013 · Authors:Erik G.F. Thomas. View a PDF of the paper titled A polarization identity for multilinear maps, by Erik G.F. Thomas. View PDF. Abstract ...Missing: homogeneous polynomials
  29. [29]
    [PDF] CLASSICAL INVARIANT THEORY
    . (c) The full polarization Pf of a homogeneous polynomial f ∈ K[V ] of degree d which was defined in 4.4 has the following description: Pf = ∆d0 ···∆20 ...
  30. [30]
    [PDF] Symmetric multilinear forms and polarization of polynomials
    Jan 9, 2009 · The goal of this paper is to investigate generalizations of the three con- cepts (quadratic form, homogeneous quadratic polynomial and symmetric.
  31. [31]
    On polarizations in invariant theory - ScienceDirect.com
    Given a reductive algebraic group G and a finite dimensional algebraic G-module V, we study how close is the algebra of G-invariant polynomials on V ⊕ n to ...
  32. [32]
    Lagrange's Theory of Quadratic Forms: From Algebraic Equations to ...
    It was Lagrange who sought to produce a general theory of quadratic forms, after Euler had published a number of deep and provocative studies of many ...Missing: 1759 | Show results with:1759
  33. [33]
    [PDF] Quadratic Forms and Their Applications - School of Mathematics
    These are the proceedings of the conference on “Quadratic Forms And. Their Applications” which was held at University College Dublin from 5th to.
  34. [34]
    The Curious History of Vectors and Tensors - SIAM.org
    Sep 3, 2024 · The idea of a vector as a mathematical object in its own right first appeared as part of William Rowan Hamilton's theory of quaternions.