Fact-checked by Grok 2 weeks ago

Shift matrix

In linear algebra, a shift matrix is a square featuring ones exclusively along either the superdiagonal (for the upper shift) or the subdiagonal (for the lower shift), with zeros in all other positions. This structure represents a basic operator that shifts the components of a along a chain of basis vectors without wrap-around. For instance, the 3×3 upper shift matrix S is given by S = \begin{pmatrix} 0 & 1 & 0 \\ 0 & 0 & 1 \\ 0 & 0 & 0 \end{pmatrix}, which maps the standard basis vector e_1 to zero and shifts e_2 to e_1, e_3 to e_2. The lower shift matrix, denoted Z_n for n, has ones on the subdiagonal, such as Z_4 = \begin{pmatrix} 0 & 0 & 0 & 0 \\ 1 & 0 & 0 & 0 \\ 0 & 1 & 0 & 0 \\ 0 & 0 & 1 & 0 \end{pmatrix}, and serves as the of the upper shift in certain contexts. Shift matrices are , meaning that for an n \times n , raising it to the n-th power yields the (S^n = 0), while S^{n-1} \neq 0, highlighting their role in studying matrix powers and indices of nilpotency. They form the simplest Jordan blocks for the eigenvalue zero in the Jordan , providing insight into the structure of non-diagonalizable matrices. Additionally, shift matrices commute only with specific forms of matrices, such as those constant along diagonals parallel to the , which constrains the centralizer in the matrix algebra. Beyond pure theory, shift matrices underpin applications in and . Shift matrices generate Toeplitz matrices when combined with diagonal matrices via expressions, and related cyclic variants facilitate efficient computations like the via the algorithm. In , weighted variants of shift matrices model unilateral shifts on Hilbert spaces, influencing studies in and .

Definition

Finite-dimensional shift matrices

In finite-dimensional linear algebra, the upper shift matrix U_n is an n \times n matrix defined by its entries (U_n)_{i,j} = \delta_{i+1,j} for i = 1, \dots, n-1 and j = 1, \dots, n, where \delta denotes the Kronecker delta; this places ones on the superdiagonal and zeros elsewhere. The lower shift matrix L_n is similarly defined as an n \times n matrix with entries (L_n)_{i,j} = \delta_{i,j+1} for i = 1, \dots, n and j = 1, \dots, n-1, resulting in ones on the subdiagonal and zeros elsewhere. Both matrices consist exclusively of entries that are either 0 or 1. The lower shift matrix satisfies the relation L_n = U_n^T, where ^T denotes the transpose. For example, in the case n=2, U_2 = \begin{pmatrix} 0 & 1 \\ 0 & 0 \end{pmatrix}, \quad L_2 = \begin{pmatrix} 0 & 0 \\ 1 & 0 \end{pmatrix}, and transposing U_2 yields L_2. For n=3, U_3 = \begin{pmatrix} 0 & 1 & 0 \\ 0 & 0 & 1 \\ 0 & 0 & 0 \end{pmatrix}, \quad L_3 = \begin{pmatrix} 0 & 0 & 0 \\ 1 & 0 & 0 \\ 0 & 1 & 0 \end{pmatrix}, again illustrating the transpose property. These matrices are nilpotent and thus have all eigenvalues equal to zero, with further spectral details addressed elsewhere.

Infinite-dimensional shift operators

In infinite-dimensional Hilbert spaces, the concept of shift operators extends the finite-dimensional shift matrices to operators acting on sequence spaces such as \ell^2(\mathbb{N}) and \ell^2(\mathbb{Z}). The unilateral forward shift operator S on the Hilbert space \ell^2(\mathbb{N}), where \mathbb{N} = \{1, 2, 3, \dots\}, is defined by its action on a sequence x = (x_1, x_2, x_3, \dots) as (S x)_1 = 0 and (S x)_n = x_{n-1} for n \geq 2. This operator shifts the sequence to the right, inserting a zero at the first position. The adjoint operator S^*, known as the backward shift, acts as (S^* x)_n = x_{n+1} for all n \geq 1, effectively removing the first component and shifting the rest leftward. With respect to the standard orthonormal basis \{e_k\}_{k=1}^\infty of \ell^2(\mathbb{N}), where e_k has a 1 in the k-th position and zeros elsewhere (so e_1 = (1, 0, 0, \dots)), the unilateral shift satisfies S e_k = e_{k+1} for each k \geq 1. Unlike its finite-dimensional counterparts, which are nilpotent (powers eventually yield the zero matrix), the infinite-dimensional unilateral shift S is not nilpotent because the space has no finite dimension to cause truncation. Moreover, S is an isometry, preserving norms via \|S x\| = \|x\| for all x \in \ell^2(\mathbb{N}), but it is not unitary since its range is the orthogonal complement of \operatorname{span}\{e_1\}, making it non-surjective. The bilateral shift operator U on \ell^2(\mathbb{Z}), the space of square-summable bi-infinite sequences indexed by all integers, is defined by (U x)_n = x_{n-1} for all n \in \mathbb{Z}. In terms of the orthonormal basis \{e_k\}_{k \in \mathbb{Z}} (with e_k having a 1 at position k and zeros elsewhere), this corresponds to U e_k = e_{k+1}. As a bilateral extension, U is both an and surjective, rendering it a on \ell^2(\mathbb{Z}). These infinite-dimensional shift operators were introduced in functional analysis by John von Neumann during the 1930s, initially in the context of studying extensions of symmetric operators and later connected to analyses in Hardy spaces.

Properties

Algebraic properties

The finite-dimensional shift matrices, often denoted as the upper shift matrix U_n (with 1's on the superdiagonal) and the lower shift matrix L_n (with 1's on the subdiagonal), exhibit several key algebraic properties stemming from their nilpotent structure. A fundamental relation is that the transpose of the upper shift matrix equals the lower shift matrix, and vice versa: L_n = U_n^T and U_n = L_n^T. To see this, note that transposing U_n, which has entries (U_n)_{i,j} = \delta_{i,j-1} for i,j = 1, \dots, n (where \delta is the Kronecker delta), yields (U_n^T)_{i,j} = (U_n)_{j,i} = \delta_{j,i-1} = \delta_{i,j+1}, which is precisely the form of L_n with entries on the subdiagonal. Both U_n and L_n are nilpotent matrices. Specifically, the power U_n^k for k = 1, \dots, n-1 is a matrix with 1's on the k-th superdiagonal and zeros elsewhere, while U_n^n = 0 (the zero matrix); an analogous description holds for powers of L_n, with 1's on the k-th subdiagonal. The nilpotency index of both matrices is n, meaning n is the smallest positive integer such that the n-th power is zero. This nilpotency implies that the minimal polynomial of U_n (and similarly for L_n) is x^n = 0. The rank of U_n is n-1, reflecting the single linear dependence among its columns (or rows), and more generally, \operatorname{rank}(U_n^k) = n - k for k = 1, \dots, n-1, with \operatorname{rank}(U_n^n) = 0; the same ranks hold for powers of L_n. This decreasing sequence underscores the collapse of the under repeated application. Products of these shift matrices yield idempotent matrices, which satisfy M^2 = M. In particular, both L_n U_n and U_n L_n are idempotent. For example, U_n L_n takes the explicit form of a with 0 in the (n,n) entry and 1's elsewhere, consisting of diagonal blocks that project onto the leading coordinates. The same idempotence holds for L_n U_n, which has 0 in the (1,1) entry and 1's elsewhere. These properties highlight the role of shift matrices as elementary components in matrix decompositions.

Spectral properties

The of the finite-dimensional U_n, the n \times n upper with on the superdiagonal and zeros elsewhere, consists solely of the eigenvalue \lambda = 0 with algebraic multiplicity n. This follows from its role as the for the \lambda^n, whose is \det(\lambda I - U_n) = \lambda^n. Consequently, the of U_n is 0, reflecting the absence of nonzero diagonal entries, and the is 0, confirming its noninvertibility. The geometric multiplicity of the eigenvalue 0 is 1, so the dimension of the of U_n is \dim \ker(U_n) = 1, spanned by the first vector e_1 = (1, 0, \dots, 0)^T. For the lower shift matrix L_n, with ones on the subdiagonal, the kernel is instead spanned by e_n = (0, \dots, 0, 1)^T. More generally, the kernels of powers satisfy \dim \ker(U_n^k) = k for k = 1, 2, \dots, n, illustrating the progressive growth of the null space up to the full . The canonical form of U_n is a single Jordan block of size n with eigenvalue 0 on the diagonal and ones on the superdiagonal; in standard conventions, U_n itself realizes this form. A achieving this, when needed for basis adjustments, can be effected by a lower with ones on the diagonal. In the infinite-dimensional setting, the of the unilateral on \ell^2(\mathbb{N}) is the closed unit disk \{\lambda \in \mathbb{C} : |\lambda| \leq 1\}, comprising no eigenvalues (empty point spectrum) but including continuous spectrum on the unit circle and residual spectrum in the open disk.

Construction and examples

Explicit matrix forms

The upper shift matrix, also known as the forward shift matrix, is a square with ones on the superdiagonal and zeros elsewhere. For dimension 2, it takes the form \begin{pmatrix} 0 & 1 \\ 0 & 0 \end{pmatrix}. The lower shift matrix, or backward shift matrix, has ones on the subdiagonal and zeros elsewhere; its 2×2 form is \begin{pmatrix} 0 & 0 \\ 1 & 0 \end{pmatrix}. In dimension 3, the upper shift matrix has ones at positions (1,2) and (2,3), yielding \begin{pmatrix} 0 & 1 & 0 \\ 0 & 0 & 1 \\ 0 & 0 & 0 \end{pmatrix}, while the lower shift matrix has ones at (2,1) and (3,2), giving \begin{pmatrix} 0 & 0 & 0 \\ 1 & 0 & 0 \\ 0 & 1 & 0 \end{pmatrix}. The general pattern for an n×n upper shift matrix places ones along the superdiagonal, with all other entries zero. For n=4, this is explicitly \begin{pmatrix} 0 & 1 & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 1 \\ 0 & 0 & 0 & 0 \end{pmatrix}. Shift matrices are binary (0,1)-matrices that resemble permutation matrices in their sparse structure with a single band of ones but are non-invertible due to their nilpotency. For example, squaring the 3×3 upper shift matrix produces a matrix with a single 1 at position (1,3) and zeros elsewhere, demonstrating the nilpotent shift property. In computational tools like MATLAB, an n×n upper shift matrix can be generated efficiently using the command diag(ones(n-1,1),1), which places ones on the first superdiagonal of an n×n zero matrix. The lower shift matrix follows analogously with diag(ones(n-1,1),-1).

Action on vectors

The shift matrix U_n, often referred to as the right shift (or forward shift) in finite-dimensional spaces, acts on the standard basis vectors \{e_1, \dots, e_n\} of \mathbb{R}^n by mapping U_n e_j = e_{j+1} for j = 1, \dots, n-1 and U_n e_n = 0, effectively moving each basis vector to the next position while sending the last to the zero vector. Similarly, the left shift matrix L_n satisfies L_n e_j = e_{j-1} for j = 2, \dots, n and L_n e_1 = 0, shifting basis vectors toward the first position and annihilating e_1. For a general vector x = (x_1, x_2, \dots, x_n)^T \in \mathbb{R}^n, the action of U_n produces U_n x = (0, x_1, x_2, \dots, x_{n-1})^T, inserting a zero at the leading position and truncating the last component, which corresponds to a rightward shift with zero-padding. In contrast, L_n x = (x_2, x_3, \dots, x_n, 0)^T, discarding the first component and appending a zero at the end, representing a leftward shift with zero-padding. These transformations preserve , as they are defined componentwise through , and illustrate how shift matrices embed sequential displacements within the structure. Iterative application of U_n shifts the vector rightward by k positions after k multiplications, with components beyond the n-th position lost to truncation, such that U_n^k x has leading zeros in the first k entries (or is the zero vector if k \geq n). The same holds for L_n, shifting leftward and appending zeros, reflecting the nilpotent nature of these operators where repeated action eventually yields the zero vector. As linear transformations, shift matrices represent a form of companion shift in the standard basis, akin to shearing operations that reorder coordinates without scaling, though they differ from permutation matrices by introducing loss through zero-padding rather than cyclic reordering. For example, with n=3 and x = (1, 2, 3)^T, the computation yields U_3 x = (0, 1, 2)^T, demonstrating the explicit rightward displacement.

Applications

In linear algebra and matrix theory

Shift matrices serve as canonical examples of nilpotent matrices in linear algebra, specifically representing a single block of size n for the eigenvalue 0. The n \times n U_n, defined with 1's on the superdiagonal and 0's elsewhere, satisfies U_n^n = 0 but U_n^{n-1} \neq 0, making its nilpotency index n. Any is similar to a block-diagonal matrix whose blocks are shift matrices of various sizes, underscoring their uniqueness in the . The shift matrix corresponds directly to the Frobenius of the p(x) = x^n. For a general p(x) = x^n + a_{n-1} x^{n-1} + \cdots + a_1 x + a_0, the features the shift structure in its first n-1 rows (shifting the vectors) and the negated coefficients -a_0, \dots, -a_{n-1} in the last row. When all coefficients a_i = 0, this reduces to the nilpotent U_n, which has characteristic and minimal polynomials both equal to x^n. This relation highlights the shift matrix's role in theoretical constructions for roots and rational canonical forms. The matrix exponential of the shift matrix, \exp(t U_n), provides a finite-dimensional model for shift flows. Since U_n is nilpotent, the exponential truncates after n terms in its Taylor series: \exp(t U_n) = I + t U_n + \frac{t^2}{2!} U_n^2 + \cdots + \frac{t^{n-1}}{(n-1)!} U_n^{n-1}, yielding a lower triangular matrix (in the transpose convention) with entries \frac{t^{j-i}}{(j-i)!} for j \geq i along the subdiagonals and 0's above. This structure approximates the continuous shift semigroup in delay differential equations, where the solution operator shifts the function history by time t, and finite-dimensional discretizations employ such nilpotent generators./3%3A_Systems_of_ODEs/3.8%3A_Matrix_exponentials) Powers of the shift matrix U_n^k perform k-position shifts on basis vectors, with the last k components zeroed out. For k \ll n, these powers match the action of the corresponding powers of the matrix (which generates circulant matrices), providing an to cyclic shifts in large dimensions. Combinations of powers, via polynomials in U_n, thus approximate circulant matrices, particularly useful in finite fields or where boundary truncation effects can be controlled or ignored. Historically, matrices were introduced by James Joseph Sylvester around 1850, with shift matrices emerging as key examples of nilpotents in early studies. Their systematic use in canonical forms was advanced by Camille Jordan in the late 19th century, and 20th-century texts expanded their applications in decompositions and operator theory.

In dynamical systems and ergodic theory

In dynamical systems, the infinite bilateral shift serves as a foundational model for the Bernoulli shift, defined as a measure-preserving transformation on the space \{0,1\}^\mathbb{Z} equipped with the product measure \mu, where each coordinate is independently assigned 0 or 1 with equal probability $1/2. The shift map \sigma: \{0,1\}^\mathbb{Z} \to \{0,1\}^\mathbb{Z} acts by \sigma(\theta)_n = \theta_{n+1} for all n \in \mathbb{Z}, preserving the measure \mu because the preimage of any cylinder set under \sigma has the same measure as the set itself, due to the independent product structure. This transformation is ergodic with respect to \mu, meaning that any invariant set has measure 0 or 1, which follows from the independence of the coordinates and the fact that the only invariant events are trivial under the product measure. Shift spaces, particularly subshifts of finite type (SFTs), extend this framework by modeling constrained dynamics via adjacency matrices that encode allowed transitions. An SFT is a closed, shift-invariant subset of the full shift over a finite alphabet, defined by forbidding a finite set of words; equivalently, it arises as the edge shift over a directed graph with adjacency matrix A, where sequences correspond to paths in the graph such that the terminal vertex of one edge matches the initial vertex of the next. For instance, the golden mean shift, with forbidden word {11}, has adjacency matrix A = \begin{pmatrix} 1 & 1 \\ 1 & 0 \end{pmatrix}, restricting sequences to avoid two consecutive 1s. The one-sided shift on these spaces captures topological dynamics, while the bilateral version aligns with measure-theoretic aspects. The Gauss map provides another connection, where shift operators model the dynamics of s for irrational numbers. Defined as G(x) = \{1/x\} for x \in (0,1], the map [G](/page/G) extracts the continued fraction digits of x = [a_0; a_1, a_2, \dots] via its itinerary with respect to the \{ (1/(n+1), 1/n] \}_{n=1}^\infty, such that [G](/page/G)(x) = [a_1; a_2, \dots], effectively acting as a left shift on the digit sequence. Iterates of [G](/page/G) thus simulate the symbolic dynamics of irrational rotations on the circle, approximating quadratic irrationals like the golden (\sqrt{5}-1)/2 = [0;1,1,1,\dots], a fixed point of [G](/page/G). A key invariant in these systems is , which quantifies the growth rate of admissible sequences; for the full r-shift over an r-symbol , it equals \log r, so the full 2-shift has \log 2. In SFTs, the is \log \lambda_A, where \lambda_A is the Perron-Frobenius eigenvalue of the A. This measure captures the exponential complexity of orbits, distinguishing non-conjugate shifts. In , shift matrices underpin convolutional codes, where linear feedback s process input bits to generate error-correcting outputs. The encoder uses a shift register of fixed length (constraint length), with tap vectors defining bits as linear combinations of stored bits; for example, nonsystematic nonrecursive codes employ two tap vectors to produce output streams resilient to noise via Viterbi decoding on the trellis graph. These shifts enable burst error correction in communication channels by constraining codewords to valid state transitions. Since the 1970s, shift maps have influenced through , with Roy Adler's work on Markov partitions for hyperbolic systems yielding SFT models for complex behaviors. Adler and Weiss (1970) showed how toral automorphisms induce SFTs, providing a symbolic encoding for orbits. Later, Adler and Marcus (1979) classified irreducible SFTs using and periodic points, facilitating analysis of attractors in dissipative systems.

Connections to permutation and circulant matrices

Shift matrices differ fundamentally from in their action on the vectors. A P corresponding to a full cycle satisfies P \mathbf{e}_i = \mathbf{e}_{i+1 \mod n} for the vectors \{\mathbf{e}_i\}, making it bijective and invertible, as it represents a rearrangement without loss of information. In contrast, the standard nilpotent shift matrix S, with 1's on the superdiagonal and zeros elsewhere, acts as S \mathbf{e}_i = \mathbf{e}_{i+1} for i < n but maps \mathbf{e}_n to the zero vector, rendering it non-bijective and defective with n-1. This "defective" nature arises because S truncates the shift at the boundary, preventing full invertibility unlike true . For a concrete illustration, consider the 3×3 case. The cyclic permutation matrix for a full cycle is P = \begin{pmatrix} 0 & 0 & 1 \\ 1 & 0 & 0 \\ 0 & 1 & 0 \end{pmatrix}, which satisfies P^3 = I and is invertible. The corresponding shift matrix is S = \begin{pmatrix} 0 & 1 & 0 \\ 0 & 0 & 1 \\ 0 & 0 & 0 \end{pmatrix}, with S^3 = 0, highlighting its nilpotency and lack of invertibility. Circulant matrices connect to shift matrices through generation via powers of a cyclic shift operator, but the finite nilpotent shift provides only an approximation due to truncation. A circulant matrix C is expressed as C = \sum_{k=0}^{n-1} c_k S^k, where S here denotes the invertible cyclic shift permutation matrix, ensuring the structure wraps around periodically. For the nilpotent shift matrix, powers beyond n-1 vanish, so finite shifts approximate circulants by truncating higher terms, useful in asymptotic analyses of Toeplitz forms. In group theory, shift matrices emerge as limits of signed matrices, particularly through nega-shift constructions. A nega-shift matrix is a signed matrix N of order n, with structure incorporating a -1 in the wrap-around entry, generating a group of order $2n via cyclic shifts with sign flips. Combinatorially, the shift matrix S serves as the of a directed with vertices $1 \to 2 \to \cdots \to n, where the superdiagonal 1's represent edges. The entry (S^k)_{i,j} = 1 if there is a directed of length exactly k from i to j (i.e., j = i + k \leq n), and 0 otherwise, thus counting such paths in the chain structure. This interpretation underscores the matrix's role in enumerating constrained walks without cycles, distinguishing it from the bijective paths enabled by matrices.

Role in Jordan canonical form

The shift matrix plays a central role in the Jordan canonical form, particularly for nilpotent operators. The n \times n upper shift matrix U_n, defined with 1's on the superdiagonal and 0's elsewhere, is identical to the Jordan block J_n(0) associated with the eigenvalue 0. This structure embodies the canonical representation of a single Jordan block, where the matrix shifts basis vectors in a chain: U_n e_1 = 0 and U_n e_{i+1} = e_i for i = 1, \dots, n-1, with \{e_1, \dots, e_n\} the . Any A consisting of a single block of size n is similar to U_n via an invertible change-of-basis P, satisfying P^{-1} A P = U_n. This reveals the underlying cyclic structure of the operator, transforming it into the standard shift form. For a general , the is a block-diagonal comprising multiple such shift matrices arranged along the diagonal, with block sizes determined by the operator's invariant factors. The existence of this form follows from the fact that every linear transformation on a finite-dimensional is similar to a of right shift operators. The structure of these shift blocks is determined by the dimensions of the generalized eigenspaces, specifically \dim \ker(A^k) for k = 1, 2, \dots. For a A, these dimensions match those of the corresponding shift matrix configuration: the number of Jordan blocks (shift components) equals \dim \ker(A), and the number of blocks of size at least j is \dim \ker(A^j) - \dim \ker(A^{j-1}). This uniquely identifies the sizes and multiplicities of the shift blocks, enabling the reconstruction of the full form without explicit similarity computation. In applications, the shift structure facilitates solving linear systems A x = b where A is . Transforming to the basis yields a block-upper-triangular with shift blocks, solvable via back-substitution: within each block, components are determined iteratively from the end of the chain, propagating solutions backward. For multiple blocks, the diagonal arrangement allows independent resolution per block. Algorithmically, this shift-like reduction underpins methods for computing the form, such as constructing chains from bases to build the similarity P, addressing the challenge of determining block structures through successive computations.

References

  1. [1]
    [PDF] 18.06 Linear Algebra, Problem set 4 solutions - MIT OpenCourseWare
    3 is one-dimensional, it gives a basis. Section 3.5. Problem 37: If AS = SA for the shift matrix S, show that A must have this special ... 18.06 Linear Algebra.
  2. [2]
    [PDF] A relation between some special centro-skew, near-Toeplitz ...
    1991 Mathematics Subject Classification. ... superdiagonal. For example, when n := 5 we have ... Let Zn denote the lower shift matrix which is the matrix that has.
  3. [3]
  4. [4]
    [PDF] Transformations of Matrix Equations Based on Tensor Product ...
    upper shift matrix and lower shift matrix are defined by. Un ... The upper shift matrix represents a linear transformation that shifts components of a.
  5. [5]
    [PDF] Shift Operators - ISI Bangalore
    Jun 4, 2014 · The shift operator, S, on l2 is defined as S({an}∞ n=0) = {0,a0,a1, ...}. On H2, the shift is Mz(f) = zf.
  6. [6]
    [PDF] The spectra of the unilateral shift and its adjoint - Jordan Bell
    Apr 3, 2014 · This note discusses shift operators, standard in operator theory, and their properties and spectra, based on Halmos' work.
  7. [7]
    [PDF] conv
    bilateral shift, like the unilateral one, is an isometry. Since the range of the bilateral shift is the entire space H, it is even unitary. = If en is the ...<|separator|>
  8. [8]
    [PDF] The Shift Operator
    A shift operator is a bounded linear transformation where Sen = en+1 for all n _ 0, and it is a fundamental building block in operator structure theory.
  9. [9]
    [PDF] Jordan Canonical Form of a Nilpotent Matrix Math 422 Schur's ...
    The Jordan Canonical Form (JCF) of a nilpotent matrix is the Jordan form where blocks are distributed along the diagonal in order of decreasing size.
  10. [10]
    Nilpotent matrix - StatLect
    Learn how nilpotent matrices and operators are defined. Discover their properties and how they are used. With detailed explanations, proofs and solved ...
  11. [11]
    [PDF] Matrix Characterizations of Riordan Arrays
    matrix M by an upper shift matrix U results in the entries of M being shifted upward by ... lower shift matrix. Thus, UUT = I and UT U = I − diag(1, 0, 0 ...
  12. [12]
    [PDF] Numerical Matrix Analysis - Ilse Ipsen
    (ii) What can you say about the rank of a nilpotent matrix, and the rank of an ... shift matrix . . . . . . . . . . . . . . . . . . . . 13 circular ...
  13. [13]
    [PDF] MATHEMATICS 217 NOTES - Math (Princeton)
    The characteristic polynomial of an n×n matrix A is the polynomial χA(λ) ... Proof: The Jordan block is similar to the companion matrix for the poly-.
  14. [14]
    [PDF] 48 Linear Algebra and Matrix Analysis Finite Dimensional Spectral ...
    R ecall that any nilpotent operator is a direct sum of shift operators in an appropriate basis, so the matrix i is similar to a direct sum of shift matrices.
  15. [15]
    [PDF] Matrix equation representation of the convolution equation ... - arXiv
    Oct 24, 2023 · The matrices 𝑈𝑛 and 𝐿𝑛 are called an upper shift matrix and a lower shift matrix respectively. The upper and lower shift matrices ...
  16. [16]
    [PDF] D. Matrix Powers and Exponentials
    where Sk is the shift matrix (C.10). Powers of J. ; k/ can be ... In Section D.2 we found an explicit expression for powers of a Jordan block, and we see.
  17. [17]
    [PDF] Linear Algebra and It's Applications by Gilbert Strang
    Page 1. Page 2. Linear Algebra and Its Applications. Fourth Edition. Gilbert Strang ... shift matrix B from R4 back to R3, transforming. (x1,x2,x3,x4) to (x2,x3, ...
  18. [18]
    [PDF] Realm of Matrices - Indian Academy of Sciences
    In general, any triangular matrix with zeros along the main diagonal is nilpotent. ... shift matrix. (possibly of different sizes). The above theorem is a special ...
  19. [19]
    Matrix Reference Manual: Special Matrices - Imperial College London
    A shift matrix, or lower shift matrix, Z, is a matrix with ones ... ZT has ones above the main diagonal and zeros elsewhere and is an upper shift matrix.
  20. [20]
    Finite-dimensional approximations of the shift operator - MathOverflow
    Nov 19, 2016 · The eigenvalues of that matrix are the nth roots of unity. So as n grows, the spectrum fills the unit circle (it does not fill the unit disk ...Modern developments in finite-dimensional linear algebraSpectral theorem from Jordan decomposition in infinite dimensionsMore results from mathoverflow.net
  21. [21]
    LU decomposition - StatLect
    A square matrix is said to have an LU decomposition (or LU factorization) if it can be written as the product of a lower triangular (L) and an upper triangular ...Missing: shift | Show results with:shift
  22. [22]
    [PDF] shifts and ergodic theory - UChicago Math
    We define measure-preserving and ergodic maps, and show that Bernoulli schemes are both measure-preserving and ergodic. We then define subshifts of finite type ...
  23. [23]
    [PDF] Symbolic dynamics and subshifts of finite type
    In short,. SFT's are associated to labeled directed graphs through the edge shift construction, and labeled directed graphs are linked with non-negative ...
  24. [24]
    [PDF] 3 Gauss map and continued fractions
    To find the continued fraction expansion of a number, we will exploit the relation with the symbolic coding of the Gauss map, in the same way that binary ...Missing: operators | Show results with:operators
  25. [25]
    [PDF] Entropy of subordinate shift spaces - arXiv
    Feb 9, 2015 · Every finite shift space has entropy zero. Example. The full A-shift has entropy log |A|. In particular, the full r-shift has entropy log r.Missing: degree | Show results with:degree
  26. [26]
    [PDF] Convolutional Codes and Turbo Codes
    We generate a transmission with a convolutional code by putting a source stream through a linear filter. This filter makes use of a shift register, linear.<|separator|>
  27. [27]
    Symbolic dynamics - Scholarpedia
    Nov 16, 2008 · For example, Markov partitions for hyperbolic toral automorphisms give rise to shifts of finite type, defined below (Adler and Weiss [1970])).
  28. [28]
    [PDF] arXiv:2504.07295v1 [cond-mat.stat-mech] 9 Apr 2025
    Apr 9, 2025 · The permutation matrix representation (PMR) is a special ... This matrix is also known as the shift matrix and is one of the generators of the ...
  29. [29]
    [PDF] Toeplitz and Circulant Matrices: A review - Stanford University
    A matrix of this form is called a circulant matrix. Circulant matrices arise, for example, in applications involving the discrete Fourier trans- form (DFT) and ...
  30. [30]
    [PDF] The Magic of Permutation Matrices - arXiv
    Jul 19, 2010 · matrices in the context of permutation matrices are found by the shifts of the identity matrix. That is, for order-4: P10 =... 0 1 0 ...
  31. [31]
    [PDF] Quasi-balanced weighing matrices, signed strongly regular ... - arXiv
    Feb 3, 2022 · A negacirculant matrix of order 2n is a polynomial in the signed permutation ... by the nega-shift matrix of order 2n−1, and let t and n be ...
  32. [32]
    [PDF] Graph Signal Processing: Overview, Challenges, and Applications
    Apr 3, 2023 · A graph interpretation for the DSP concepts of Section II-A can be achieved by viewing the 0-1 shift matrix Ac of (7) as the adjacency matrix of ...
  33. [33]
    [PDF] Discrete Signal Processing on Graphs: Frequency Analysis
    shift-invariant graph filters. In particular, we can define the normalized graph shift matrix. Anorm = 1. |λmax|. A,. (6) where λmax denotes the eigenvalue of A ...
  34. [34]
    [PDF] The Jordan Canonical Form of a Nilpotent Matrix Math 422 Schur's ...
    Dec 7, 2010 · Since the diagonal entries of D are the eigenvalues of L; and. = 0 is the only eigenvalue of L, we have. D = 0: Solving PL1LP = 0 for L gives L ...
  35. [35]
    The Jordan normal form and the Euclidean algorithm - Terry Tao
    Oct 12, 2007 · For instance, every right shift operator is nilpotent, as is any direct sum of right shifts. ... Matrix) which has no Eigenvalues ne. 0 ...
  36. [36]
    [PDF] The Weyr Characteristic - Swarthmore College
    Jan 12, 1999 · The list of the non-increasingly ordered sizes of the diagonal blocks corresponding to an eigenvalue a is called the Weyr characteristic of. A ...