Fact-checked by Grok 2 weeks ago

Semi-orthogonal matrix

In linear algebra, a semi-orthogonal is a rectangular with real entries whose columns (or rows, depending on the dimensions) form an orthonormal set, satisfying either A^T A = I_n for an m \times n with m \geq n or A A^T = I_m for m < n. This property ensures that the acts as a partial isometry, preserving the Euclidean norm of vectors in the appropriate subspace. Such matrices are fundamental in matrix decompositions, including the , where a full-rank m \times n matrix A (with m \geq n) can be factored as A = QR, with Q semi-orthogonal and R upper triangular. They also appear in the of rectangular matrices, expressing A = PV where P is semi-orthogonal and V is positive semidefinite. For symmetric matrices, semi-orthogonal matrices facilitate spectral decompositions by providing orthonormal bases for eigenspaces of reduced dimension. Beyond decompositions, semi-orthogonal matrices find applications in numerical linear algebra and statistics, such as in least squares estimation and principal component analysis, where they enable efficient orthogonal projections. In signal processing, they model phenomena like turbulent fluctuations by maintaining orthogonality in subspaces. Recent research explores additional constraints, such as when semi-orthogonal matrices have row vectors of equal lengths, which occurs under specific scaling conditions and relates to Grassmannian coordinates for column spaces. These properties distinguish semi-orthogonal matrices from fully orthogonal square matrices, which satisfy both A^T A = I and A A^T = I.

Definition and Terminology

Formal Definition

A semi-orthogonal matrix is a rectangular real matrix Q \in \mathbb{R}^{m \times n} whose entries satisfy specific orthonormality conditions depending on the dimensions m and n. When m \geq n (a tall matrix), the columns of Q form an orthonormal set, satisfying Q^\top Q = I_n, where I_n denotes the n \times n identity matrix; in this case, Q preserves the Euclidean norm for all vectors x \in \mathbb{R}^n, i.e., \| Q x \|_2 = \| x \|_2. When m \leq n (a short matrix), the rows of Q form an orthonormal set, satisfying Q Q^\top = I_m, where I_m denotes the m \times m identity matrix; here, Q preserves the Euclidean norm for all vectors x in the row space of Q, i.e., \| Q x \|_2 = \| x \|_2 for x in the row space of Q. The prefix "semi-" highlights the partial orthogonality arising from the non-square shape, in contrast to a full orthogonal matrix, which is square and satisfies both Q^\top Q = Q Q^\top = I (or the analogous condition for unitary matrices over the complex numbers).

Equivalent Characterizations

A semi-orthogonal matrix can be characterized as a partial isometry, meaning it maps vectors in the orthogonal complement of its kernel to vectors of the same Euclidean norm in the codomain. Specifically, for a matrix Q \in \mathbb{R}^{m \times n}, \| Qx \|_2 = \| x \|_2 holds for all x such that x \perp \ker(Q). This property extends the notion of an isometry to rectangular matrices, where the isometric action is restricted to a subspace of the domain. In the tall case where m \geq n and Q^\top Q = I_n, Q serves as a sub-isometry by preserving the Euclidean norm for all vectors in the entire domain \mathbb{R}^n, as the kernel is trivial under full column rank. In the short case where m \leq n and Q Q^\top = I_m, Q preserves norms on its row space, which is the orthogonal complement of the kernel. These cases highlight how semi-orthogonal matrices embed isometric mappings into higher- or lower-dimensional spaces. An equivalent matrix-theoretic characterization involves the singular values: for a semi-orthogonal matrix Q, the singular values satisfy \sigma_i(Q) = 1 for i = 1, \dots, \min(m,n), with any remaining singular values being zero if the matrix is not square. This follows from the fact that the non-zero singular values correspond to the unit eigenvalues of Q^\top Q or Q Q^\top. Unlike full isometries, which are square preserving norms across the entire domain and codomain, semi-orthogonal matrices act as isometries only on appropriate subspaces, allowing for dimension mismatch while maintaining partial norm preservation.

Core Properties

Orthonormality Conditions

A semi-orthogonal matrix satisfies specific orthonormality conditions that distinguish it from fully orthogonal matrices, applying to either its columns or rows depending on the matrix dimensions. For a tall semi-orthogonal matrix Q \in \mathbb{R}^{m \times n} with m > n, the columns \mathbf{q}_i and \mathbf{q}_j (for i, j = 1, \dots, n) are orthonormal, meaning their inner products satisfy \mathbf{q}_i^\top \mathbf{q}_j = \delta_{ij}, where \delta_{ij} is the (equal to 1 if i = j and 0 otherwise). This condition ensures that the columns form an for the column space of Q. Similarly, for a short semi-orthogonal matrix Q \in \mathbb{R}^{m \times n} with m < n, the rows \mathbf{r}_i and \mathbf{r}_j (for i, j = 1, \dots, m) satisfy \mathbf{r}_i \mathbf{r}_j^\top = \delta_{ij}, establishing orthonormality among the rows. These pairwise orthonormality requirements extend to a preservation of inner products within the appropriate subspace. For the tall case, the transformation Q preserves the Euclidean inner product for vectors \mathbf{x}, \mathbf{y} \in \mathbb{R}^n, such that \langle Q \mathbf{x}, Q \mathbf{y} \rangle = \langle \mathbf{x}, \mathbf{y} \rangle. In the short case, this preservation holds for vectors in the row space, aligning with the orthonormality of the rows. As a direct consequence, the relevant equals the identity on the corresponding dimensions: Q^\top Q = I_n for tall matrices and Q Q^\top = I_m for short matrices. This property underscores the role of semi-orthogonal matrices in forming idempotent projections, where Q Q^\top (for tall) or Q^\top Q (for short) acts as an orthogonal projector onto the column or row space, respectively.

Norm Preservation

A key property of semi-orthogonal matrices is their preservation of the Euclidean norm on specific subspaces, distinguishing them as partial isometries in linear transformations. For a tall semi-orthogonal matrix [Q](/page/Q) \in \mathbb{R}^{m \times n} with m > n and satisfying Q^\top Q = I_n, the Euclidean norm is preserved under the mapping [Q](/page/Q): \mathbb{R}^n \to \mathbb{R}^m, meaning \| [Q](/page/Q) x \|_2 = \| x \|_2 for all x \in \mathbb{R}^n. This norm preservation arises directly from the orthonormality condition on the columns of Q. A brief sketch of the derivation shows that \| Q x \|_2^2 = x^\top Q^\top Q x = x^\top I_n x = x^\top x = \| x \|_2^2, confirming the equality of norms without altering lengths. For a short semi-orthogonal matrix Q \in \mathbb{R}^{m \times n} with m < n and satisfying Q Q^\top = I_m, the preservation holds on the row space of Q, so \| Q x \|_2 = \| x \|_2 for all x in the row space of Q. Geometrically, the columns of a tall Q (or rows of a short Q) form an orthonormal frame for the subspace, ensuring that vectors are embedded or projected without distortion in length, akin to a rigid rotation or reflection within that frame. In contrast to general matrices, which typically distort the Euclidean norm through scaling, shearing, or contraction/expansion, semi-orthogonal matrices induce rigid transformations that maintain distances on their defined subspaces.

Full Rank and Singular Values

A semi-orthogonal matrix Q \in \mathbb{R}^{m \times n} has full rank equal to \min(m, n), as its columns (or rows, depending on the orientation) form an orthonormal set and are thus linearly independent. The singular values of Q consist of exactly \min(m, n) values equal to 1 and the remaining |m - n| values equal to 0, reflecting its structure as a partial isometry. For the case where m \geq n (tall matrix), the eigenvalues of Q^\top Q are all 1 (with multiplicity n), confirming the non-zero singular values are 1. Similarly, for m < n (short matrix), the eigenvalues of Q Q^\top are all 1 (with multiplicity m). This singular value structure implies that a tall semi-orthogonal matrix Q (with m \geq n) admits a left inverse given by Q^\top, since Q^\top Q = I_n. Conversely, a short semi-orthogonal matrix admits a right inverse Q^\top, as Q Q^\top = I_m. In contrast to fully orthogonal square matrices, where all singular values are exactly 1, semi-orthogonal matrices exhibit a partial spectrum of 1's due to their rectangular nature.

Examples

Tall Matrices

A tall semi-orthogonal matrix features orthonormal columns, providing an isometric embedding from \mathbb{R}^n into \mathbb{R}^m for m > n. A basic example is the $2 \times 1 matrix Q = \begin{pmatrix} \frac{1}{\sqrt{2}} \\ \frac{1}{\sqrt{2}} \end{pmatrix}, whose single column is a , satisfying Q^\top Q = 1. Another example arises in the Householder process, where the thin Q factor is a tall with orthonormal columns; for instance, consider the $3 \times 2 Q = \begin{pmatrix} \frac{1}{\sqrt{3}} & 0 \\ \frac{1}{\sqrt{3}} & -\frac{1}{\sqrt{2}} \\ \frac{1}{\sqrt{3}} & \frac{1}{\sqrt{2}} \end{pmatrix}, with columns that are unit vectors and mutually orthogonal, as Q^\top Q = I_2. Geometrically, the columns of this Q span a plane in \mathbb{R}^3, and left-multiplication by Q maps vectors from \mathbb{R}^2 into \mathbb{R}^3 without distorting lengths or angles within that plane. To illustrate norm preservation, for the first example with input x = 1, we have Qx = Q and \|Qx\|_2 = 1 = \|x\|_2; similarly, for the second example with x = \begin{pmatrix} 1 \\ 0 \end{pmatrix}, Qx is the first column of Q with \|Qx\|_2 = 1 = \|x\|_2, and for x = \begin{pmatrix} 0 \\ 1 \end{pmatrix}, Qx is the second column with matching norms.

Short Matrices

A short semi-orthogonal is an m \times n with m < n whose rows form an orthonormal set in \mathbb{R}^n. A basic example is the $1 \times 2
Q = \begin{pmatrix} \frac{1}{\sqrt{2}} & \frac{1}{\sqrt{2}} \end{pmatrix}.
The single row has Euclidean norm \sqrt{ \left( \frac{1}{\sqrt{2}} \right)^2 + \left( \frac{1}{\sqrt{2}} \right)^2 } = \sqrt{ \frac{1}{2} + \frac{1}{2} } = 1, so Q Q^T = 1, verifying row-orthonormality.
For a $2 \times 3 example from coordinate averaging, consider rows derived by normalizing averages of coordinate directions, such as the first row from averaging the first and second standard basis vectors in \mathbb{R}^3:
\mathbf{r}_1 = \frac{1}{\sqrt{2}} ( \mathbf{e}_1 + \mathbf{e}_2 ) = \begin{pmatrix} \frac{1}{\sqrt{2}} & \frac{1}{\sqrt{2}} & 0 \end{pmatrix},
and the second row orthogonal to it and unit length, obtained via on an average involving the third coordinate, such as
\mathbf{r}_2 = \frac{1}{\sqrt{2}} ( \mathbf{e}_1 - \mathbf{e}_2 ) = \begin{pmatrix} \frac{1}{\sqrt{2}} & -\frac{1}{\sqrt{2}} & 0 \end{pmatrix}.
The matrix is
Q = \begin{pmatrix} \frac{1}{\sqrt{2}} & \frac{1}{\sqrt{2}} & 0 \\ \frac{1}{\sqrt{2}} & -\frac{1}{\sqrt{2}} & 0 \end{pmatrix}.
Each row has norm 1, and their dot product is \frac{1}{2} - \frac{1}{2} + 0 = 0, so Q Q^T = I_2. This spans the xy-plane subspace.
Such matrices enable isometric projections of \mathbb{R}^3 onto \mathbb{R}^2 within the row space, embedding the parameter space isometrically via Q^T. To check norm preservation on row space vectors, take y = \begin{pmatrix} 1 \\ 0 \end{pmatrix} \in \mathbb{R}^2. Then Q^T y = \begin{pmatrix} \frac{1}{\sqrt{2}} \\ \frac{1}{\sqrt{2}} \\ 0 \end{pmatrix} \in \mathbb{R}^3, with \| Q^T y \| = 1 = \| y \|. Applying Q recovers Q (Q^T y) = y, preserving the norm. A similar check holds for y = \begin{pmatrix} 0 \\ 1 \end{pmatrix}, yielding Q^T y = \begin{pmatrix} \frac{1}{\sqrt{2}} \\ -\frac{1}{\sqrt{2}} \\ 0 \end{pmatrix} with norm 1.

Non-Examples

A semi-orthogonal matrix requires that for a tall matrix (more rows than columns), the columns are orthonormal such that Q^\top Q = I, or for a short matrix (more columns than rows), the rows are orthonormal such that Q Q^\top = I. Matrices failing these conditions serve as non-examples, highlighting the necessity of proper normalization and orthogonality. Consider a tall 2×1 matrix Q = \begin{pmatrix} 1 \\ 1 \end{pmatrix}. Here, Q^\top Q = [1, 1] \begin{pmatrix} 1 \\ 1 \end{pmatrix} = 2 \neq 1, so the column is not of unit length, violating the semi-orthogonality condition for tall matrices. For a short non-example, take the 1×2 matrix Q = [1, 1]. Then Q Q^\top = [1, 1] \begin{pmatrix} 1 \\ 1 \end{pmatrix} = 2 \neq 1, meaning the row lacks unit length and fails the orthonormality requirement for short matrices. In the square case, semi-orthogonality coincides with full orthogonality, requiring both Q^\top Q = I and Q Q^\top = I. A counterexample is the 2×2 shear matrix Q = \begin{pmatrix} 1 & 1 \\ 0 & 1 \end{pmatrix}, where Q Q^\top = \begin{pmatrix} 1 & 1 \\ 1 & 2 \end{pmatrix} \neq I_2, as the columns are neither orthogonal nor unit length. A common pitfall involves matrices that appear normalized in one direction but lack full rank, such as the 2×2 zero matrix Q = \begin{pmatrix} 0 & 0 \\ 0 & 0 \end{pmatrix}. Here, Q^\top Q = \begin{pmatrix} 0 & 0 \\ 0 & 0 \end{pmatrix} \neq I_2, failing the orthonormality condition due to zero singular values and absence of a full orthonormal basis.

Proofs

Preservation of Euclidean Norm

A semi-orthogonal matrix preserves the of vectors under its linear transformation, a direct consequence of its orthonormality conditions. Consider first the tall case, where Q \in \mathbb{R}^{m \times n} with m \geq n and Q^\top Q = I_n. For any vector x \in \mathbb{R}^n, the squared Euclidean norm satisfies \| Q x \|^2 = (Q x)^\top (Q x) = x^\top Q^\top Q x = x^\top I_n x = x^\top x = \| x \|^2. Thus, \| Q x \| = \| x \| for all x \in \mathbb{R}^n. This preservation holds because the columns of Q form an orthonormal basis for the column space. In the short case, where Q \in \mathbb{R}^{m \times n} with m < n and Q Q^\top = I_m, the situation differs. Here, Q^\top Q is the orthogonal projection onto the row space of Q, a subspace of dimension m in \mathbb{R}^n. In general, \| Q x \|^2 = x^\top Q^\top Q x \leq \| x \|^2, making Q non-expansive. However, norm preservation occurs when x lies in the row space of Q, i.e., x = Q^\top z for some z \in \mathbb{R}^m. In this case, Q x = Q Q^\top z = I_m z = z, \quad \| Q x \| = \| z \|, and \| x \|^2 = (Q^\top z)^\top (Q^\top z) = z^\top Q Q^\top z = z^\top I_m z = \| z \|^2 = \| Q x \|^2. Thus, \| Q x \| = \| x \| for x in the row space. This norm preservation extends to inner products. For the tall case, \langle Q x, Q y \rangle = x^\top Q^\top Q y = x^\top y = \langle x, y \rangle for all x, y \in \mathbb{R}^n. Similarly, in the short case, \langle Q x, Q y \rangle = x^\top Q^\top Q y = x^\top y = \langle x, y \rangle when both x and y are in the row space of Q. These properties follow analogously from the respective orthonormality assumptions.

Full Column or Row Rank

A semi-orthogonal matrix Q \in \mathbb{R}^{m \times n} with m \geq n satisfies Q^\top Q = I_n, meaning its columns form an orthonormal set. This orthonormality ensures the columns are linearly independent. To prove this, suppose \sum_{i=1}^n a_i \mathbf{q}_i = \mathbf{0} for some coefficients a_1, \dots, a_n \in \mathbb{R}, where \mathbf{q}_i denotes the i-th column of Q. Taking the inner product with \mathbf{q}_j yields a_j = \mathbf{q}_j^\top \left( \sum_{i=1}^n a_i \mathbf{q}_i \right) = \sum_{i=1}^n a_i (\mathbf{q}_j^\top \mathbf{q}_i) = \sum_{i=1}^n a_i \delta_{ji} = a_j, but since the left side is zero, a_j = 0 for all j. Thus, the only solution is the trivial one, confirming linear independence and full column rank n = \min(m, n). Equivalently, the condition Q^\top Q = I_n implies the kernel of Q is trivial: if Q \mathbf{x} = \mathbf{0} for \mathbf{x} \in \mathbb{R}^n, then \mathbf{x}^\top Q^\top Q \mathbf{x} = \mathbf{x}^\top \mathbf{x} = 0, so \mathbf{x} = \mathbf{0}. This injectivity further establishes full column rank. By the contrapositive, if \operatorname{rank}(Q) < n, then \operatorname{rank}(Q^\top Q) < n, but Q^\top Q = I_n has full rank n, a contradiction, since \operatorname{rank}(Q^\top Q) = \operatorname{rank}(Q). For the case m \leq n, a semi-orthogonal matrix Q satisfies Q Q^\top = I_m, implying its rows form an orthonormal set and are linearly independent, yielding full row rank m = \min(m, n). The proof mirrors the column case: suppose \sum_{i=1}^m b_i \mathbf{r}_i = \mathbf{0}^\top for row vectors \mathbf{r}_i, then taking the inner product with \mathbf{r}_j gives b_j = 0 for all j. Similarly, \operatorname{rank}(Q Q^\top) = \operatorname{rank}(Q) = m, confirming the result.

Singular Value Constraints

The singular value decomposition (SVD) of an m \times n real matrix Q expresses it as Q = U \Sigma V^T, where U is an m \times m orthogonal matrix, V is an n \times n orthogonal matrix, and \Sigma is an m \times n diagonal matrix containing the singular values \sigma_1 \geq \sigma_2 \geq \cdots \geq \sigma_{\min(m,n)} \geq 0 along its main diagonal. The singular values \sigma_i are the square roots of the eigenvalues of Q^T Q (or equivalently, of Q Q^T). For a tall semi-orthogonal matrix Q (with m \geq n and Q^T Q = I_n), the matrix Q^T Q is the n \times n identity matrix, whose eigenvalues are all equal to 1. Thus, the first n singular values satisfy \sigma_i = \sqrt{1} = 1 for i = 1, \dots, n, and any remaining singular values (if m > n) are zero. For a short semi-orthogonal matrix Q (with n \geq m and Q Q^T = I_m), the matrix Q Q^T is the m \times m , whose eigenvalues are all equal to 1. Thus, the first m singular values satisfy \sigma_i = 1 for i = 1, \dots, m, and any remaining singular values (if n > m) are zero. This singular value structure implies that the Frobenius norm of Q is \|Q\|_F = \sqrt{\sum_{i=1}^{\min(m,n)} \sigma_i^2} = \sqrt{\min(m,n)}, since there are exactly \min(m,n) non-zero singular values each equal to 1.

Applications and Relations

In QR Decomposition

In the QR decomposition of a full-rank m \times n matrix A with m \geq n, the factorization takes the form A = QR, where Q is an m \times n semi-orthogonal matrix with orthonormal columns satisfying Q^T Q = I_n, and R is an n \times n upper triangular matrix. This thin or reduced QR form is particularly useful for matrices where the number of rows exceeds the number of columns, ensuring computational efficiency by avoiding the full m \times m orthogonal matrix. One common method to construct this decomposition is the Gram-Schmidt process, which sequentially orthogonalizes the columns of A to produce the semi-orthogonal Q, with the upper triangular R capturing the coefficients from the orthogonalization steps. The classical requires approximately $2mn^2 floating-point operations, though it can suffer from numerical instability in finite precision arithmetic due to subtractive cancellation. A modified version improves stability by performing projections in a more careful order, making it suitable for many practical computations. For enhanced , especially in dense cases, the Householder QR algorithm employs a series of orthogonal reflections to triangularize A, resulting in a semi-orthogonal Q that can be represented compactly via the Householder vectors stored in the lower triangular part of the modified A. This approach requires about $2n^2 (m - n/3) operations and is backward stable, meaning the computed accurately reflects a nearby to the original. Alternatively, Givens rotations can be used to zero out elements selectively, producing the same semi-orthogonal Q, though at a higher cost of roughly $3n^2 (m - n/3) operations; this method excels in sparse or structured matrices where only specific entries need modification. A primary application of semi-orthogonal Q in arises in solving overdetermined problems, \min_x \| Ax - b \|_2, for tall full-rank A. After computing A = QR, the problem reduces to solving the triangular system R x = Q^T b, where the of Q's columns preserves the Euclidean norm, ensuring \| Ax - b \|_2 = \| R x - Q^T b \|_2 and thus minimizing the residual efficiently without forming the normal equations. This approach avoids the ill-conditioning often associated with normal equations, leveraging the stability of the QR process for reliable solutions in .

Relation to Other Matrices

A tall semi-orthogonal matrix Q \in \mathbb{R}^{m \times n} with m > n and orthonormal columns satisfies Q^T Q = I_n, making it an that preserves the Euclidean norm on its domain, i.e., \|Q x\|_2 = \|x\|_2 for all x \in \mathbb{R}^n. This property positions Q as an isometric from \mathbb{R}^n into \mathbb{R}^m. Furthermore, the product QQ^T forms the orthogonal onto the column space of Q. In (), semi-orthogonal matrices represent the principal directions, where the loading matrix V (of size p \times k, k < p) has orthonormal columns, allowing dimension reduction while maximizing variance preservation in the projected subspace. Semi-orthogonal matrices also appear in the of rectangular matrices, where a full-rank m \times n matrix A (with m \geq n) factors as A = U P, with U semi-orthogonal (orthonormal columns) and P . This decomposition highlights the isometric and metric-preserving aspects of semi-orthogonal matrices. In the case of a short semi-orthogonal matrix Q \in \mathbb{R}^{m \times n} with m < n and orthonormal rows, Q Q^T = I_m, rendering Q a co-isometry, as it is a surjective preserving norms on the . The transpose Q^T then becomes a tall semi-orthogonal matrix. Such matrices relate to frame theory in , where semi-orthogonal frames provide redundant representations with orthogonal components across scales, as seen in semi-orthogonal frame wavelets for multiresolution analysis. They also model transitions from stable flows to turbulent fluctuations in signals, such as electrocardiogram patterns. Recent research (as of ) explores constraints where semi-orthogonal matrices have row vectors of equal lengths, achievable via row under no accidental linear relations, relating to coordinates for column spaces and the matrix problem. Semi-orthogonal matrices parametrize the V_{m,n}, the set of all m \times n matrices with orthonormal columns when m \geq n, which is diffeomorphic to the quotient O(m)/O(m-n). In the complex domain, the analog is the semi-unitary matrix, satisfying Q^H Q = I_n or Q Q^H = I_m, extending the condition to Hermitian inner products and appearing in tensor decompositions like the canonical polyadic decomposition for . More broadly, semi-orthogonal matrices are finite-dimensional instances of partial isometries in , where the operator preserves norms on the of its kernel.

References

  1. [1]
    None
    Below is a merged response that consolidates all the information from the provided summaries into a single, comprehensive summary. To maximize detail and clarity, I will use a table in CSV format to organize the key information (Definition, Page Number, Context, and Useful URLs) across the different segments. Following the table, I will provide a narrative summary to tie everything together and address any discrepancies or notes.
  2. [2]
    [PDF] On Orthogonalities in Matrices - arXiv
    The third type is called as semi-orthogonal matrix, which is defined by Abadir and Magnus in. [1], as follows: Definition 2.3. If A is a real m × n matrix ...
  3. [3]
    [PDF] On semi-orthogonal matrices with row vectors of equal lengths
    Nov 13, 2024 · Abstract. When does a rectangular matrix with an orthonormal set of column vectors have row vectors of equal lengths?
  4. [4]
    Matrix Algebra - Cambridge University Press & Assessment
    Abadir, University of York, Jan R. Magnus, Universiteit van Tilburg. Karim M ... Cambridge University Press. Publication date: June 2012: August 2005. ISBN ...
  5. [5]
    [PDF] Partially isometric matrices
    Sep 27, 2025 · The square matrix A is a partial isometry if and only if A = QE and if and only if. A = FQ where Q is unitary, E and F are orthogonal.Missing: semi- | Show results with:semi-
  6. [6]
    [PDF] Unit II: Numerical Linear Algebra Chapter II.3: QR Factorization, SVD
    Equivalently, QT Q = QQT = I. Orthogonal matrices preserve the Euclidean norm of a vector, i.e.. kQvk2. 2 = vT QT Qv = vT v = kvk2. 2. Hence, geometrically, we ...
  7. [7]
    None
    ### Summary of Semi-Orthogonal Matrices from Lecture 05
  8. [8]
    [PDF] partially isometric matrices: a brief and selective survey
    These ma- trices are characterized by the equation AA∗A = A, in which A∗ denotes the con- jugate transpose of A. We refer to such a matrix as a partial isometry ...
  9. [9]
    The Singular Value Decomposition - JMP
    ... semi-orthogonal matrix with V′V = Ir. Diag(Λ) is an r by r diagonal matrix with positive diagonal elements given by the column vector Λ = (λ1 ... matrix ...
  10. [10]
    [PDF] Orthogonal Matrices and the Singular Value Decomposition
    Even in the general case, the singular values of a matrix A are the lengths of the semi-axes of the hyperellipse E defined by. E = {Ax : kxk = 1} . The SVD ...
  11. [11]
    What are the singular values of an orthogonal matrix? What ... - Quora
    Nov 18, 2016 · Therefore, all singular values of an orthogonal matrix are equal to 1 1 . To see this without invoking the decomposition, recall that the ...Why are the singular values of a matrix (from SVD) called so ... - QuoraHow unique are $u$ and $v$ in the singular value decomposition ...More results from www.quora.com
  12. [12]
    [PDF] Orthogonal matrices and Gram-Schmidt - MIT OpenCourseWare
    Matrices with orthonormal columns are a new class of important matri ces to add to those on our list: triangular, diagonal, permutation, symmetric, reduced row ...
  13. [13]
    [PDF] Householder QR - CS@Cornell
    Householder QR uses Householder transformations, which are simple orthogonal transformations corresponding to reflection through a plane, to compute QR ...
  14. [14]
    [PDF] 5. Orthogonal matrices
    ... • complex matrices with orthonormal columns. Page 17. Tall matrix with orthonormal columns suppose 𝐴 ∈ R𝑚×𝑛 is tall (𝑚>𝑛) and has orthonormal columns.
  15. [15]
    None
    ### Definition and Examples of Semi-Orthogonal Matrix
  16. [16]
  17. [17]
    [PDF] Golub and Van Loan - EE IIT Bombay
    Finally, we mention that the 2-norm is preserved under orthogonal transformation. ... The SVD is an orthogonal matrix reduction and so the 2-norm and Frobenius ...
  18. [18]
    [PDF] Orthogonalizing Convolutional Layers with the Cayley Transform
    Apr 14, 2021 · norm-preserving semi-orthogonal matrix; the case cin ≥ cout follows similarly, but the resulting matrix is merely non-expansive. With ...
  19. [19]
    [PDF] Linear Algebra - UC Berkeley Statistics
    Jul 28, 2006 · You should be able now to prove the following results: 1. rank(A0A) = rank(A). In particular, if rank(A) = n (columns are linearly independent) ...
  20. [20]
    [PDF] Singular Value Decomposition of Real Matrices
    Mar 13, 2020 · The diagonal entries of Σ are called the singular values of A. The column vectors of V are called the right singular vectors of A.
  21. [21]
    None
    Below is a merged summary of the QR Decomposition sections from "Matrix Computations" (4th Ed.), consolidating all information from the provided summaries into a comprehensive response. To retain maximum detail, I will use a structured format with tables where appropriate, followed by a narrative summary for additional context. The response avoids any thinking tokens and focuses on presenting the information as requested.
  22. [22]
    [PDF] Linear Algebra - Math 121B - UCI Mathematics
    If T is a unitary/orthogonal operator, then it is a linear isometry. 2. If T is a linear isometry and V is finite-dimensional, then T is unitary/orthogonal.
  23. [23]
    [PDF] Stiefel-Grassmann-manifolds-Edelman.pdf - CIS UPenn
    Points in the Grassmann manifold are equivalence classes of n-by-p orthogonal matrices, where two matrices are equivalent if their columns span the same p- ...
  24. [24]
    [PDF] CS 3220: Orthogonal projectors - CS@Cornell
    Specifically, given a matrix V ∈ Rn×k with orthonormal columns. P = V V T is the orthogonal projector onto its column space. An alternative way of saying ...
  25. [25]
    Semi-orthogonal Frame Wavelets and Parseval Frame ... - NIH
    In this paper, we study semi-orthogonal frame wavelets and Parseval frame wavelets(PFWs) in L 2 (R d ) with matrix dilations of form ( D f ) ( x ) = 2 f ( A x
  26. [26]
    [PDF] Canonical Polyadic decomposition with a Columnwise Orthonormal ...
    Jan 25, 2013 · For convenience, we say that this matrix is semi-unitary (in the complex case) or semi-orthogonal (in the real case). The CPD of a higher ...<|separator|>
  27. [27]
    Partial isometries and pseudoinverses in semi-Hilbertian spaces
    In this article the concepts of partial isometries, normal partial isometries and generalized projections in the context of operators defined on a semi- ...