Fact-checked by Grok 2 weeks ago

Ultrametric space

An ultrametric space is a set equipped with a that satisfies the strong : for all points x, y, z in the space, the d(x, z) \leq \max\{d(x, y), d(y, z)\}. This condition strengthens the standard of spaces and implies that all triangles are isosceles, with each side no longer than the longest of the other two. Ultrametric spaces exhibit distinctive topological properties, such as the equality of open and closed balls, the nesting of balls (where intersecting balls have one contained in the other), and total disconnectedness, meaning no non-trivial connected subsets exist. These spaces are non-Archimedean and often model hierarchical structures, leading to applications in diverse s including p-adic analysis, where the p-adic numbers \mathbb{Q}_p form a complete ultrametric under the p-adic valuation. Notable examples include the discrete metric on any set, where distances are 1 between distinct points and 0 otherwise; the space of sequences with the metric d(a, b) = \rho^n (for $0 < \rho < 1), with n the first differing index; and the Cantor set endowed with an ultrametric derived from its ternary expansion. In biology and statistics, ultrametrics arise in codon spaces and hierarchical clustering methods like ascending hierarchical classification. The concept was introduced by Marc Krasner in 1944, motivated by valuations on fields, with further developments by Jean-Pierre Benzécri in 1972 linking it to statistical analysis. Ultrametric spaces have since influenced areas such as fractal geometry, disordered systems in physics, and optimization in random media.

Fundamentals

Definition

An ultrametric space is a set X together with a function d: X \times X \to [0, \infty) that satisfies the following axioms for all x, y, z \in X:
  • d(x, y) = 0 if and only if x = y,
  • d(x, y) = d(y, x),
  • d(x, z) \leq \max\{d(x, y), d(y, z)\}.
This third axiom, known as the or , replaces the standard d(x, z) \leq d(x, y) + d(y, z) of a . The ultrametric inequality implies the usual one, since the maximum of two nonnegative numbers is at most their sum, making every ultrametric space a metric space but with stricter distance constraints that often lead to hierarchical or tree-like structures. The function d is termed the ultrametric distance. Such spaces are also called non-Archimedean metric spaces, a synonym reflecting their connection to non-Archimedean valuations on fields, where the Archimedean axiom—that for any positive elements a and b, there exists a natural number n such that na > b—fails. The term "non-Archimedean" stems from Ostrowski's 1917 classification of s on fields, distinguishing those satisfying the ultrametric from the Archimedean real absolute value. In an ultrametric space, open balls B(x, r) = \{y \in X \mid d(x, y) < r\} exhibit a distinctive symmetry: every point y \in B(x, r) serves as a center for the same ball, so B(x, r) = B(y, r).

Historical Development

The origins of ultrametric spaces lie in the development of non-Archimedean valuations within algebraic number theory. In 1844, Ernst Kummer introduced the concept of ideal numbers to resolve the failure of unique prime factorization in rings of integers of cyclotomic fields. This work laid important groundwork for algebraic number theory, which later incorporated non-Archimedean valuations satisfying the ultrametric inequality to measure the "size" of elements in a way that strengthens the triangle inequality to a maximum condition. The formalization of ultrametric spaces advanced significantly with Kurt Hensel's introduction of p-adic numbers in 1897. In his paper "Über eine neue Begründung der Theorie der algebraischen Zahlen," Hensel constructed the p-adic completion of the rationals with respect to the p-adic valuation, yielding a field where the associated metric is ultrametric. This construction provided the first concrete example of a complete ultrametric space, enabling rigorous analysis in number theory and highlighting the topological peculiarities of such spaces, such as totally disconnected yet complete structures. The explicit term "ultrametric space" was coined by Marc Krasner in 1944, in his work "Nombres semi-réels et espaces ultramétriques," where he generalized the strong triangle inequality to abstract metric spaces motivated by p-adic constructions. This naming emphasized the "ultra" strengthening of the standard metric axiom, distinguishing it from ordinary metric spaces. In the mid-20th century, ultrametric spaces gained broader recognition through connections to functional analysis, particularly via non-Archimedean and spectral theory. Pioneering works, influenced by I. M. 's foundational contributions to abstract functional analysis and representation theory, extended these ideas to p-adic settings, as seen in developments like non-Archimedean . Initially confined to number theory for problems in Diophantine equations and algebraic geometry, the framework evolved into applications in topology, operator algebras, and beyond, with dedicated conferences on p-adic functional analysis emerging by the 1990s.

Properties

Core Axioms and Inequalities

An ultrametric space (X, d) is defined by a d that satisfies the standard axioms of positivity, symmetry, and the identity of indiscernibles, along with the strong triangle inequality: for all x, y, z \in X, d(x, z) \leq \max\{d(x, y), d(y, z)\}. This strong form replaces the usual triangle inequality d(x, z) \leq d(x, y) + d(y, z). The strong inequality implies the standard one directly, as \max\{d(x, y), d(y, z)\} \leq d(x, y) + d(y, z) holds for nonnegative reals, so d(x, z) \leq d(x, y) + d(y, z). To see this step-by-step, assume without loss of generality that d(x, y) \geq d(y, z); then d(x, z) \leq d(x, y) \leq d(x, y) + d(y, z). The case d(y, z) > d(x, y) follows symmetrically. Thus, every ultrametric space is a . A key consequence of the strong triangle inequality is the isosceles property: for any x, y, z \in X, at least two of the distances d(x, y), d(y, z), and d(z, x) are equal. To prove this, suppose for contradiction that all three distances are distinct, say d(x, y) < d(y, z) < d(z, x). By the strong applied to d(z, x), d(z, x) \leq \max\{d(z, y), d(y, x)\} = \max\{d(y, z), d(x, y)\} = d(y, z), since d(y, z) > d(x, y). But this yields d(z, x) \leq d(y, z) < d(z, x), a contradiction. The other orderings of distinct distances lead to similar contradictions by symmetry and relabeling. If exactly two distances are equal and larger than the third, the triangle is isosceles with the two equal sides as the longer ones; if all three are equal, it is equilateral. Moreover, if d(x, y) \neq d(y, z), then d(x, z) = \max\{d(x, y), d(y, z)\}, ensuring the two largest sides are equal. The strong triangle inequality also imparts a hierarchical structure to distance relations in ultrametric spaces. For points x, y, z \in X, the inequality d(x, z) \leq \max\{d(x, y), d(y, z)\} often achieves equality, particularly along chains where distances increase monotonically. Specifically, if d(x, y) \neq d(y, z), equality holds as shown above, meaning the distance "jumps" to the maximum without intermediate values. In longer chains x_1, x_2, \dots, x_n, the distance d(x_1, x_n) equals the maximum pairwise consecutive distance \max_{i=1}^{n-1} d(x_i, x_{i+1}), derived inductively: assuming it holds up to x_{k}, then d(x_1, x_{k+1}) \leq \max\{d(x_1, x_k), d(x_k, x_{k+1})\}, and equality follows if the maxima differ by the strong inequality's sharpness. This hierarchical nesting reflects tree-like geometries, where subsets are organized by distance levels without "crossing" branches. Ultrametrics are preserved under positive scalar multiplication: if d' is defined by d'(x, y) = c \cdot d(x, y) for some c > 0, then (X, d') is also an ultrametric. The positivity and axioms hold immediately, and the strong follows as d'(x, z) = c \cdot d(x, z) \leq c \cdot \max\{d(x, y), d(y, z)\} = \max\{c \cdot d(x, y), c \cdot d(y, z)\} = \max\{d'(x, y), d'(y, z)\}. Such rescalings are uniformly equivalent to the original , as the map \mathrm{id}: (X, d) \to (X, d') is bi-Lipschitz with constants c and $1/c, ensuring the same uniform structure and . A fundamental states that in an ultrametric space, the of any closed is at most its . Consider the closed ball B(x, r) = \{y \in X \mid d(x, y) \leq r\}. For any y, z \in B(x, r), d(y, z) \leq \max\{d(y, x), d(x, z)\} \leq r, so the \mathrm{diam}(B(x, r)) = \sup\{d(y, z) \mid y, z \in B(x, r)\} \leq r. Moreover, let s = \sup\{d(x, y) \mid y \in B(x, r)\} \leq r. Then \mathrm{diam}(B(x, r)) = s, since d(y, z) \leq \max\{d(y, x), d(x, z)\} \leq s and, taking y = x, \sup\{d(x, z) \mid z \in B(x, r)\} = s. Thus, the equals the supremum of distances from the center, which is at most r. This property underscores the "uniform spread" within balls, contrasting with general metric spaces where diameters can reach up to twice the radius.

Topological and Metric Characteristics

In an ultrametric space (X, d), the topology is generated by the d, where the collection of all open balls B(x, r) = \{ y \in X \mid d(x, y) < r \} for x \in X and r > 0 forms a basis for the . This basis ensures that the space is Hausdorff and first-countable. A defining feature of the ultrametric is that every open ball is also closed, making all balls clopen sets. To see this, consider the complement of an open ball B(x, r); any point y \notin B(x, r) satisfies d(x, y) \geq r, and by the strong , for any z \in B(x, r), d(y, z) = d(y, x) \geq r since d(x, z) < r implies d(y, z) = d(y, x) \geq r. Thus, B(y, r) is disjoint from B(x, r), showing B(x, r) is closed. The clopen nature of balls implies the space admits a basis of clopen sets, rendering it zero-dimensional. As a consequence, ultrametric spaces are totally disconnected: the only connected subsets are singletons, since any two distinct points can be separated by disjoint clopen balls. Ultrametric spaces are paracompact, meaning every open cover admits a locally finite open refinement. This follows from the metric structure, where the clopen basis allows for refinements by shrinking balls to avoid overlaps beyond finite collections in compact neighborhoods; specifically, ultrametric spaces are strictly paracompact, as the strong triangle inequality ensures nested balls permit precise control over covers. Regarding completeness, an ultrametric space is complete if every Cauchy sequence converges to a point in the space. In such spaces, the strong triangle inequality implies that Cauchy sequences have unique limits: if (x_n) is Cauchy, then for large n, m, d(x_n, x_m) \to 0, and the limit x satisfies d(x_n, x) = \sup_{k \geq n} d(x_k, x) decreasing to zero monotonically. A criterion for completeness is that the space is spherically complete, meaning every decreasing sequence of non-empty closed balls with radii tending to zero has non-empty intersection; this is equivalent to every Cauchy sequence converging, as the terms eventually lie in such nested balls. The metric completion of any ultrametric space remains ultrametric, preserving these properties. Metric characteristics of ultrametric spaces often manifest in a tree-like structure, where points can be hierarchically organized via nested balls resembling branches of a tree, with distances corresponding to the lowest common ancestor in the tree metric. This structure ensures no cycles in the "distance graph," as the ultrametric inequality enforces that paths between points branch without loops, modeling hierarchical clustering without circular dependencies. Ultrametric spaces extend naturally to uniform spaces via the uniformity generated by the entourages \{(x, y) \mid d(x, y) < \epsilon\} for \epsilon > 0, which is compatible with the and satisfies the uniform continuity properties strengthened by the ultrametric condition. In complete uniform ultrametric spaces, the holds: the space is of second category, as countable intersections of dense open sets remain , with adaptations leveraging the clopen basis for non-Archimedean uniformity to prove without relying on sequential .

Examples

Discrete and Finite Examples

A singleton set equipped with the metric d(x, x) = 0 forms a trivial ultrametric space, as the only point satisfies all axioms vacuously. On any X, the defined by d(x, y) = 1 if x \neq y and d(x, x) = 0 satisfies the ultrametric inequality, since for distinct points the maximum of distances is 1, matching the direct distance, while equal points yield 0. Finite ultrametric spaces arise naturally from the leaves of a rooted where all root-to-leaf paths have equal length; the distance between two leaves is twice the distance from their to the leaves, ensuring the strong triangle inequality holds as the maximum pairwise distance equals the direct one. For strings of equal length over a finite , an ultrametric can be defined by d(s, t) = \rho^k, where $0 < \rho < 1 and k is the length of the longest common prefix (i.e., the position of the first differing symbol, or n if identical), which satisfies the ultrametric property by taking the minimum k (maximum distance) among paths. In the specific case of binary strings in \{0,1\}^n, a 2-adic-like valuation induces an ultrametric by interpreting strings as approximations to 2-adic integers modulo $2^n, with distance d(x, y) = 2^{-v_2(x - y)} where v_2 is the 2-adic valuation of the difference (highest power of 2 dividing it, corresponding to the highest differing bit position); this structure embeds the finite space into the 2-adic metric while preserving ultrametricity. A modification of the Hamming distance on strings over a finite alphabet, such as weighting differences by their ultrametric depth in a hierarchical clustering, yields an ultrametric when the underlying structure enforces tree-like separations, as seen in genetic sequence analysis where codon distances respect maximal pairwise separations.

Analytic and Infinite Examples

One prominent example of an infinite ultrametric space arises from any field equipped with a non-Archimedean absolute value, which induces a metric satisfying the ultrametric inequality d(x,y) \leq \max\{d(x,z), d(z,y)\} for all x,y,z in the field. Such valued fields, where the absolute value |\cdot| satisfies |x+y| \leq \max\{|x|,|y|\}, yield ultrametric topologies that are totally disconnected and complete under appropriate completions. The field of p-adic numbers \mathbb{Q}_p, for a prime p, exemplifies this construction as the completion of the rational numbers \mathbb{Q} with respect to the p-adic valuation v_p, defined by v_p(a/b) = v_p(a) - v_p(b) where v_p(n) counts the highest power of p dividing the integer n. The associated metric is d_p(x,y) = p^{-v_p(x-y)}, which is non-Archimedean and renders \mathbb{Q}_p a complete ultrametric space, enabling series expansions analogous to real Taylor series but with respect to powers of p. Elements of \mathbb{Q}_p can be represented as formal series \sum_{k=n}^\infty a_k p^k with digits a_k \in \{0,1,\dots,p-1\} and n \in \mathbb{Z}, converging in this metric. Formal power series fields provide another analytic construction, such as k[], the ring of formal power series over a field k in the indeterminate t, equipped with the t-adic valuation v_t(f) = \min\{n \mid a_n \neq 0\} for f = \sum_{n=0}^\infty a_n t^n. The induced metric d(f,g) = q^{-v_t(f-g)} for some q > 1 (often q=2 if k has characteristic not 2) defines an ultrametric on the field of fractions k((t)), which consists of Laurent series \sum_{n=m}^\infty a_n t^n with m \in \mathbb{Z}, making it complete and locally compact. This space models function fields in positive characteristic and supports rigid analytic geometry over non-Archimedean bases. The space of rational functions over a non-Archimedean valued inherits an ultrametric from non-Archimedean , such as the Gauss norm on k(t) where for polynomials p = \sum a_i t^i, \|p\| = \max_i |a_i|, extended to quotients by \|f/g\| = \|f\| / \|g\|. A geometric realization of an ultrametric space appears in the classical , which can be endowed with an ultrametric via the 3-adic valuation on expansions restricted to digits 0 and 2, interpreting points as limits of sequences avoiding the middle-third intervals. Specifically, the distance between two points in the C \subset [0,1] is d(x,y) = 3^{-k} where k is the largest integer such that the expansions of x and y agree on the first k digits (using 0 and 2), inducing a non-Archimedean that aligns with the hierarchical of the set's construction. This embedding highlights how fractal-like sets can carry ultrametric topologies, facilitating analysis of measure growth and diffusion processes on C.

Applications

In Number Theory and Algebra

In , ultrametric spaces play a fundamental role through the classification of s on the rational numbers \mathbb{Q}. Ostrowski's theorem states that every non-trivial on \mathbb{Q} is equivalent to either the standard archimedean |\cdot|_\infty or a p-adic |\cdot|_p for some prime p, where the p-adic valuations induce ultrametric norms on the completions \mathbb{Q}_p. These p-adic s satisfy the strong |x + y|_p \leq \max(|x|_p, |y|_p), making \mathbb{Q}_p a prototypical ultrametric space that underpins much of modern . The ultrametric structure of \mathbb{Q}_p is essential for solving Diophantine equations by assessing local solubility in p-adic fields. exploits this ultrametric property to lift solutions of equations from modulo p to full solutions in the p-adic integers \mathbb{Z}_p, provided the derivative condition holds to ensure and in the non-archimedean . For instance, if a f(x) \equiv 0 \pmod{p} with f'(a) \not\equiv 0 \pmod{p}, then there exists a unique p-adic root lifting a, the verification of whether Diophantine equations have solutions in \mathbb{Q}_p as a necessary condition for global rational solutions. This local solubility check via ultrametric completions is a cornerstone for computational and theoretical approaches to Diophantine problems. In , ultrametric spaces arise as completions of global fields at non-archimedean places, forming local fields that are complete with respect to discrete valuations. For a number field K, the completion at a \mathfrak{p} yields a K_\mathfrak{p} equipped with an ultrametric , generalizing the p-adic case. These completions are integral to the adele ring \mathbb{A}_K, defined as the restricted of all K_v over places v of K, where the non-archimedean components contribute ultrametric structure to facilitate the study of global units, class groups, and zeta functions. The adele ring encodes local information globally, with its reflecting the ultrametric nature of the finite places, and supports the idelic formulation of the Artin reciprocity map in . Ultrametric local fields are crucial in local-global principles, such as the Hasse-Minkowski theorem, which asserts that a quadratic form over \mathbb{Q} represents zero non-trivially if and only if it does so over \mathbb{R} and all \mathbb{Q}_p. Here, solubility in the ultrametric spaces \mathbb{Q}_p provides the non-archimedean local conditions, leveraging the compactness of the p-adic unit ball to classify isotropic quadratic forms via Hilbert symbols that are multiplicative in the ultrametric norm. This principle extends to more general Diophantine problems, where failure of p-adic solubility obstructs global solutions, highlighting the ultrametric framework's role in bridging local and global arithmetic. Connections to Galois representations and further underscore the importance of ultrametric completions in . p-adic Galois representations, arising from of \mathrm{Gal}(\overline{\mathbb{Q}}/\mathbb{Q}) on groups H^i_{\ét}(\overline{X}, \mathbb{Q}_p), are studied within the category of representations over the ultrametric field \mathbb{Q}_p, where the non-archimedean ensures and de Rham properties. These representations link geometric objects over number fields to arithmetic data via p-adic , with ultrametric completions providing the local fields for crystalline and syntomic comparisons that refine the . Seminal results, such as Fontaine's theory, classify potentially crystalline representations using weakly admissible modules over the ultrametric ring of p-adic periods, advancing the understanding of motives and Langlands correspondences.

In Geometry and Data Analysis

In data analysis, ultrametric spaces provide a natural framework for representing hierarchical structures through dendrograms, which encode clustering hierarchies where the distance between points corresponds to the height at which their branches merge in single-linkage algorithms. This equivalence between dendrograms and ultrametric spaces allows for efficient computation of hierarchical clusters, as the ultrametric distance u(x_i, x_j) is defined as the minimum r such that x_i and x_j belong to the same at level r. Such representations are particularly useful in generative models, where dendrogram distances evaluate clustering fidelity by comparing sorted merge heights between real and generated data distributions. In , ultrametric spaces model evolutionary by assuming a constant rate of divergence, known as the hypothesis, where all leaves are equidistant from the , reflecting contemporaneous tips. Ultrametric distances thus capture the minimum bottleneck paths in evolutionary networks, enabling the construction of ultrametric phylogenetic trees that approximate divergence patterns from dissimilarity data. For instance, the space of ultrametric phylogenetic trees forms a cubical or under metrics like the \tau-space or t-space parameterizations, facilitating computations for summarizing tree ensembles in evolutionary studies. Geometrically, ultrametric spaces can be interpreted as taxicab geometries on , where distances follow paths along branching structures, and their boundaries at align with end spaces of \mathbb{R}- via categorical equivalences that preserve local isometries and similarities. This tree-like structure approximates spaces, as ultrametric spaces admit fillings by hyperbolizing nested balls, yielding Gromov 0- metrics whose boundaries quasimöbiusly map to the original space, akin to embeddings. Ultrametricity, measured via three-point conditions in hyperbolicity vectors, quantifies deviations from tree metrics and informs fitting algorithms that embed into such geometries with bounded distortion. In computer science, ultrametric embeddings accelerate nearest-neighbor searches in databases by mapping high-dimensional data to ultrametric spaces, where queries resolve in constant time via dendrogram traversals that identify the tightest containing cluster. Subquadratic algorithms achieve (1 + \epsilon)-approximations of optimal ultrametric fits, enabling efficient hierarchical representations for large-scale retrieval tasks in Õ(n^{2 - \epsilon + o(\epsilon^2)}) time. In , min-plus algebra—where addition is minimization and multiplication is —induces ultrametrics on spaces, as the tropical linear space of phylogenetic trees coincides with the of tree ultrametrics under max-plus operations. This models varieties as tropical , such as the \ell_\infty-nearest ultrametric polytope, supporting geometric optimizations like Fermat-Weber points over M-ultrametric spaces.

In Physics and Optimization

Ultrametric spaces have significant applications in the physics of disordered systems, particularly in spin glasses, where the replica symmetry breaking (RSB) scheme reveals an ultrametric structure in the . This hierarchical organization models the energy landscape of complex systems like the Sherrington-Kirkpatrick model, where overlaps between states follow ultrametric relations, facilitating the analysis of glassy dynamics and relaxation processes. In optimization problems within random media, ultrametricity emerges in directed polymers and landscape models, capturing the tree-like structure of minimal energy paths. For instance, in directed polymers in random media (DPRM), the optimal transport substates exhibit ultrametric properties, aiding the study of localization and roughness exponents in disordered environments. These models extend to high-dimensional function optimization, such as in mixed p-spin spin glasses, where ultrametric hierarchies describe the complexity of local minima and barrier crossings. Additionally, ultrametric spaces model fractal geometries with hierarchical , such as in the construction of ultrametric sets or approximations of dimensions in tree-like structures, bridging metric properties with geometric complexity.

References

  1. [1]
    Espace ultramétrique et valuation - EuDML
    Krasner, Marc. "Espace ultramétrique et valuation." Séminaire Dubreil. Algèbre et théorie des nombres 1 (1947-1948): 1-17. <http://eudml.org/doc/111045> ...
  2. [2]
    Ultrametric Space - an overview | ScienceDirect Topics
    An ultrametric space is defined as a metric space that satisfies the strong triangle inequality, which means that for any three points, the distance between ...
  3. [3]
    [PDF] An introduction to the geometry of ultrametric spaces - arXiv
    Nov 5, 2007 · The discrete metric on any set M is defined by setting d(x, y) = 1 when x 6= y. This is an ultrametric, and there are also more interesting ...
  4. [4]
  5. [5]
    Definition and Elementary Properties of Ultrametric Spaces
    An ultrametric space is a metric space in which the triangle inequality is strengthened by using the maximum instead of the sum.
  6. [6]
    [PDF] Ultrametric spaces - Kiran S. Kedlaya
    An ultrametric field, or nonarchimedean valued field (in the terminology of [FvdP]), is a field. K equipped with a function |·| : K → R≥0 with the following ...
  7. [7]
    [PDF] General Definitions
    • Ultrametric. An ultrametric (or non-Archimedean metric) is (Krasner 1944) a metric d on X which satisfies the following strengthened version of the ...
  8. [8]
    [PDF] Origin and Development of Valuation Theory - ICTS
    The classification of absolute values into archimedean and non-archimedean was given by Ostrowski in 1917. An absolute value φ is said to be non-archimedean if ...
  9. [9]
    [PDF] Ultrametrics and valuations - Assets - Cambridge University Press
    Ultrametric spaces will form the building blocks of the locally convex spaces to be treated in this book, whereas valued fields will act as their scalar fields.Missing: source | Show results with:source
  10. [10]
    Eduard Kummer (1810 - 1893) - Biography - MacTutor
    In 1843 Kummer, realising that attempts to prove Fermat's Last Theorem broke down because the unique factorisation of integers did not extend to other rings of ...
  11. [11]
    Kurt Hensel - Biography - MacTutor - University of St Andrews
    Kurt Hensel was born in East Prussia, in the city then called Königsberg. His father was Sebastian Hensel who was a landowner at the time that Kurt was born.
  12. [12]
    Earliest Known Uses of Some of the Words of Mathematics (U)
    The term ULTRAMETRIC was coined in 1944 by M. Krasner, according to an Internet web page. The terms UMBRAL CALCULUS and UMBRAL NOTATION were coined by James ...
  13. [13]
    [PDF] Path Integrals on Ultrametric Spaces - DSpace@MIT
    really date back to the invention by Kurt Hensel in 1897 of p-adic number fields. [Hen97], which have proven invaluable in number theory ever since.<|separator|>
  14. [14]
    Non-Archimedean Gelfand theory - MSP
    It is shown that under certain conditions a specified sub- algebra Xo of a non-Archimedean Banach algebra is iso- metrically isomorphic to the space of all ...
  15. [15]
    Introduction to Metric Fixed Point Theory
    Thus, a spherically complete ultrametric space is complete. Some nice and immediate consequences of isosceles property are: (a) If F is a family of closed ...
  16. [16]
    [PDF] Ultrametrics and Complete Multipartite Graphs
    Inequality (1.1) is often called the strong triangle inequality. It is clear that every ultrametric space is a metric space, but not vice versa in general.
  17. [17]
    [PDF] Urysohn ultrametric spaces and isometry groups. - UNT Digital Library
    In 1927, Urysohn [12] defined a Polish metric space which turned out to contain all other. Polish spaces as closed subspaces. Theorem 1.1 (Urysohn). There ...<|separator|>
  18. [18]
    [PDF] arXiv:2407.21763v1 [math.GN] 31 Jul 2024
    Jul 31, 2024 · As a consequence of point 5 of Proposition 2.6, we get that the topology induced by an ultrametric has a clopen basis. Thus, an ultrametrizable.Missing: "ultrametric
  19. [19]
    [PDF] arXiv:2207.12765v4 [math.MG] 10 Aug 2022
    Aug 10, 2022 · A topological space is said to be totally disconnected if every its ... ultrametric space via the. Gromov-Hausdorff ultrametric, Topology Appl.
  20. [20]
    [PDF] p-Adic Analysis and Lie Groups
    Any ultrametric space X is strictly paracompact. ii. Any compact space X is paracompact. Proof. i. This follows from Lemma 1.4. ii. This is trivial ...
  21. [21]
    [PDF] trees and ultrametric spaces: a categorical equivalence
    Oct 18, 2002 · How noncommutative geometry can be used to study the micro-geometry of ultrametric spaces and the geometry of trees at infinity constitutes the ...<|control11|><|separator|>
  22. [22]
    [PDF] Hereditary properties of finite ultrametric spaces - arXiv
    Dec 23, 2024 · Gelfand: using graph theory describe up to isometry all finite ul- trametric spaces. An appropriate representation of ultrametric spaces X by ...
  23. [23]
    [PDF] arXiv:2008.10209v2 [math.MG] 11 Mar 2021
    Mar 11, 2021 · Ultrametric space, Baire space, Isometric embedding. The author was ... The following is known as the Baire category theorem. Theorem ...
  24. [24]
    [PDF] 20.2 Ultrametric distances
    A tree is said to have additive edge lengths if the distance between two leaves is the sum of the edge lengths connecting them. You can show that ultrametric ...<|separator|>
  25. [25]
    [PDF] Ultrametrics in the genetic code and the genome - arXiv
    Apr 9, 2017 · In the case of strings as parts of DNA, RNA and proteins, this modified distance is finer and should be more appropriate than Hamming distance ...
  26. [26]
    [PDF] Summary on non-Archimedean valued fields
    Definition 1.15. An ultrametric space is called spherically complete if each nested sequence of balls has a non-empty intersection. Remark 1.16. The concept of ...
  27. [27]
    [PDF] p-ADIC ABSOLUTE VALUES - UChicago Math
    An absolute value is non-archimedean if and only if it induces an ultrametric space. Theorem 4.3 ([2, Proposition 2.3.3]). For a non-archimedean absolute value, ...
  28. [28]
    [PDF] p-adic numbers - Assets - Cambridge University Press
    A set together with a metric is called a metric space. A metric d is called non-Archimedean (or ultra-metric) if it satisfies the additional condition. (3. ) ...
  29. [29]
    [PDF] AN INTRODUCTION TO THE p-ADIC NUMBERS - UChicago Math
    Unlike the real numbers, the p-adic numbers are an ultrametric space, leading to a number of fascinating but often counterintuitive results. The p-adic ...
  30. [30]
    [PDF] Introduction to p-adic Numbers
    Dec 3, 2022 · is called an ultrametric space. Formally,. Definition: An ultrametric space is a metric space satisfying the ultrametric inequality: d(x, y) ...
  31. [31]
    [PDF] Elementary properties of power series fields over finite fields ∗
    The power series field Fp((t)), also called “field of formal Laurent series ... They can be given already in the context of ultrametric spaces, but here we will ...
  32. [32]
    Approximation orders of formal Laurent series by Oppenheim ...
    In other words, ||·|| is a non-Archimedean norm of the field . It is well known that is a complete metric space under the metric ρ defined by ρ(A,B)=||A−B||. ...
  33. [33]
    [PDF] Topology of Nonarchimedean Analytic Spaces
    Jan 11, 2013 · For instance, K could be the field of rational numbers Q, the field of formal Laurent series C((t)), the function field of an algebraic curve.
  34. [34]
    (PDF) Ultrametric Cantor sets and growth of measure - ResearchGate
    An interesting phenomenon, called growth of measure, is studied on such an ultrametric space. Using the reparametrization invariance of the valuation it is ...Missing: via | Show results with:via
  35. [35]
    [PDF] OSTROWSKI'S THEOREM The prime numbers also arise in a very ...
    Ostrowski's Theorem states that prime numbers arise as the possible ways of defining absolute values on Q, and these examples give all possible absolute values.
  36. [36]
    [PDF] 1 Absolute values on fields - Kiran S. Kedlaya
    Conversely, if |·| is a nonarchimedean absolute value, then v(·) = −log |·| is a real valuation. It is worth noting that there are comparatively few archimedean ...
  37. [37]
    [PDF] HENSEL'S LEMMA 1. Introduction In the p-adic integers ...
    Hensel's Lemma gives conditions under which a root of a polynomial mod p lifts to a root in Zp, such as the polynomial X2 − 7 with p = 3.
  38. [38]
    [PDF] the p-adics, hensel's lemma, and newton polygons - UChicago Math
    Aug 30, 2013 · This paper aims to develop the theory of p-adic integers and their basic algebraic and topological properties. It introduces the field of p-adic.
  39. [39]
    [PDF] A multivariable Hensel's lemma - Keith Conrad
    Hensel's lemma in Zp[X] is the following result about refining an approximate solution of f(X) = 0 to an exact solution. then for a unique α ∈ Zp, f(α) = 0 and ...Missing: solubility | Show results with:solubility
  40. [40]
    [PDF] Adeles and Ideles and their applications - UChicago Math
    Aug 18, 2010 · Adeles are a restricted direct product of completions of a number field, while ideles are the invertible elements within the adele ring.Missing: ultrametric | Show results with:ultrametric
  41. [41]
    [PDF] adeles in mathematical physics - arXiv
    Jul 26, 2007 · p-Adic numbers and adeles have many applications in mathematics, e.g. represen- tation theory, algebraic geometry and modern number theory.
  42. [42]
    [PDF] p-ADIC NUMBERS, HASSE-MINKOWSKI THEOREM, APPLICATIONS
    Aug 31, 2019 · Abstract. This paper discusses the topic of finding rational and integer so- lutions of an equation with homogeneous polynomials of degree 2 ...
  43. [43]
    [PDF] Hasse-Minkowski Theorem - CSUN
    Let p = pi for some i. We claim that a is a square modulo p. This is trivial if a ≡ 0 (mod p). Otherwise a is a p-adic unit, and by assumption there exists ...
  44. [44]
    [PDF] Theory of p-adic Galois Representations
    Math. Soc. 1 (1988), 255–299. [Fal89]. , Crystalline cohomology and p-adic étale cohomology, Algebraic analysis, geometry and number theory, The John Hopkins ...
  45. [45]
    [PDF] Ultrametric Information Spaces - arXiv
    Feb 15, 2012 · A convenient data structure for points in an ultrametric space is a den- drogram. We define a dendrogram as a rooted, labeled, ranked, binary ...
  46. [46]
    [PDF] Dendrogram distance: an evaluation metric for generative networks ...
    Nov 29, 2023 · Using results from clustering theory, we demonstrate the equivalence between dendrograms and ultrametric spaces to develop a metric that has ...<|separator|>
  47. [47]
    Random ultrametric trees and applications
    We show how the boundary of an ultrametric tree, like any compact ultrametric space, can be represented in a simple way via the so-called comb metric.
  48. [48]
    Ultrametric networks: a new tool for phylogenetic analysis - PMC
    Mar 5, 2013 · In this paper, we will study the ultrametric networkG(V|W) associated with W, i.e. the graph G(V|W):=(V, E(V|W)) with vertex set V and edge set ...
  49. [49]
    The space of ultrametric phylogenetic trees - ScienceDirect.com
    Aug 21, 2016 · We introduce two natural metric spaces on ultrametric phylogenetic trees. The mean (summary) tree is different for different parameterisations.
  50. [50]
    (PDF) A Hyperbolic Filling for Ultrametric Spaces - ResearchGate
    Aug 6, 2025 · In this paper we define a hyperbolic filling for a perfect ultrametric space by way of hyperbolizing the collection of its nonde-generate balls.
  51. [51]
    [PDF] Fitting trees to l1-hyperbolic distances - arXiv
    Sep 2, 2024 · An ultrametric space is one in which the usual triangle inequality has been strengthened to d(x, y) ≤ max{d(x, z),d(y, z)}. A hyperbolic metric ...
  52. [52]
    A (1 + ε)-approximation for ultrametric embedding in subquadratic time
    Aug 5, 2025 · Efficiently computing accurate representations of high-dimensional data is essential for data analysis and unsupervised learning.
  53. [53]
    Tropical Geometric Variation of Tree Shapes
    Jul 12, 2022 · The tropical linear space coincides with the space of tree ultrametrics, and hence, that tropically-convex sets (and therefore tropical line ...
  54. [54]
    Tropical Fermat-Weber Points over Spaces of 𝑀-Ultrametrics - arXiv
    Sep 20, 2025 · We view the space of M -ultrametrics through the lens of tropical geometry. The tropical max-plus semiring is 𝕋 := ( ℝ ∪ { − ∞ } , ⊕ , ⊙ ) , ...