Fact-checked by Grok 2 weeks ago

Box topology

In topology, the box topology is a standard topology defined on the Cartesian product \prod_{\alpha \in J} X_\alpha of an arbitrary indexed family of topological spaces (X_\alpha, \tau_\alpha), where the basis for this topology consists of all sets of the form \prod_{\alpha \in J} U_\alpha such that U_\alpha is an in X_\alpha for every \alpha \in J. This construction allows each component space to contribute an arbitrary to the basis elements, making the box topology particularly suited for infinite products where uniformity across all coordinates is emphasized. Unlike the coarser product topology, which restricts basis elements to those where all but finitely many U_\alpha = X_\alpha, the box topology is strictly finer when J is infinite, meaning every product-open set is box-open, but not conversely. The two topologies coincide precisely when the index set J is finite, ensuring that finite products retain the familiar structure from basic . A notable consequence is that all projection maps \pi_\beta: \prod_{\alpha \in J} X_\alpha \to X_\beta are continuous in both the product and box topologies, but the finer box topology renders certain maps—such as the diagonal map—discontinuous for infinite J that are continuous in the product topology. The box topology exhibits several distinctive properties, particularly in infinite-dimensional settings. For instance, on spaces like \mathbb{R}^\mathbb{N} (the set of all real-valued sequences), it is not metrizable, as sequences that converge often do not converge in the box topology—such as the sequence f_n defined by f_n(k) = 1/n for all k, which does not approach the zero sequence. It also preserves certain separation axioms like Hausdorffness from the component spaces but may fail to be or paracompact in infinite products. In the context of function spaces Y^X, where X is a set and Y a topological space, the box topology provides a natural framework for studying over all points, though it contrasts with the or compact-open topologies in and preservation.

Definition and Construction

Formal Definition

The product space X = \prod_{i \in I} X_i of a family of topological spaces (X_i, \tau_i), where I is an arbitrary , consists of all functions x: I \to \bigcup_{i \in I} X_i such that x(i) \in X_i for each i \in I. The projection maps are the functions \pi_j: X \to X_j defined by \pi_j(x) = x(j) for each j \in I. The box topology \tau_b on X is the topology generated by the collection \mathcal{B} = \left\{ \prod_{i \in I} U_i \;\middle|\; U_i \in \tau_i \text{ for all } i \in I \right\} as a basis. Each B \in \mathcal{B} is called a basis element of the box topology. This basis satisfies the conditions for a basis of a topology: for any B_1 = \prod_{i \in I} U_i and B_2 = \prod_{i \in I} V_i in \mathcal{B} with x \in B_1 \cap B_2, there exists B_3 = \prod_{i \in I} W_i \in \mathcal{B} such that x \in B_3 \subseteq B_1 \cap B_2, where W_i = U_i \cap V_i for each i. The box topology \tau_b is the unique topology on X that makes all projection maps \pi_j continuous and has the standard basis elements \prod_{i \in I} U_i (with each U_i open in X_i) as open sets.

Basis Elements

In the box topology on a product space X = \prod_{i \in I} X_i, where each X_i is a , the basis consists of all sets of the form \prod_{i \in I} U_i, with U_i open in X_i for every i \in I. Unlike the , there is no requirement that all but finitely many U_i equal X_i; each U_i may be a proper open independently. Every in the box topology is a of such basis elements, and the topology is the unique one generated by this collection as a basis. This collection forms a basis for a on X because it satisfies the axioms: first, the union of all basis elements covers X, as for any point (x_i)_{i \in I} \in X, one can choose open neighborhoods V_i \ni x_i in each X_i to form \prod_{i \in I} V_i containing the point. Second, the intersection of any two basis elements \prod_{i \in I} U_i and \prod_{i \in I} V_i is either empty or equal to \prod_{i \in I} (U_i \cap V_i), which is again a basis element since each U_i \cap V_i is open in X_i. Thus, every point in the intersection lies in a basis element contained within it. The basis elements themselves are open in the generated topology by construction. If each X_i has a basis \mathcal{B}_i for its , then the collection \left\{ \prod_{i \in I} B_i \mid B_i \in \mathcal{B}_i \ \forall i \in I \right\} forms a basis for the box topology on X. The box topology is the finest topology on X such that the canonical projection maps \pi_j: X \to X_j are continuous for all j \in I, as any coarser topology would fail to include some of these basis elements as open sets. For a simple example, consider the finite product \mathbb{R}^2 = \mathbb{R} \times \mathbb{R}, where the standard topology on each \mathbb{R} has basis of open intervals. A typical basis element for the box topology is (a, b) \times (c, d), an open , and these generate the usual on the . In the infinite case, such as X = \mathbb{R}^\mathbb{N} = \prod_{n=1}^\infty \mathbb{R}, a basis element takes the form \prod_{n=1}^\infty (a_n, b_n), where each (a_n, b_n) is an open interval in \mathbb{R}; this specifies an independent open interval constraint on every coordinate, yielding sets like "all sequences whose nth term lies in (a_n, b_n) for each (n$."

Core Properties

Separation and Connectedness

The box topology on a product space \prod_{i \in I} X_i preserves the basic separation axioms from the factor spaces. Specifically, the product is T_0 (Kolmogorov) if and only if each X_i is T_0, as the property relies on distinguishing points via open sets in at least one coordinate, which the continuous projections \pi_j: \prod X_i \to X_j preserve. Similarly, the product is T_1 (Fréchet) if and only if every X_i is T_1, since singletons are closed in each factor, making them closed in the box product via the full support of basis elements. For Hausdorff separation (T_2), the box product is Hausdorff precisely when each X_i is Hausdorff. To see this, suppose x \neq y in \prod X_i; then there exists some index j with x_j \neq y_j. Since X_j is Hausdorff, there are disjoint opens U_j, V_j \subset X_j containing x_j, y_j respectively. The sets U = \prod_{i \in I} U_i and V = \prod_{i \in I} V_i, where U_i = X_i = V_i for i \neq j and U_j, V_j as above, form disjoint open neighborhoods of x, y in the box topology, as they are basic open sets. The converse holds because if the product is Hausdorff, the projections, being continuous and open, map to Hausdorff images. The box product also inherits regularity (T_3) from the factors: if each X_i is regular, so is \prod X_i in the box topology. For a point x \in \prod X_i and closed C \subset \prod X_i with x \notin C, the set of coordinates where x and points of C differ allows construction of disjoint opens using regularity in those factors and full products of opens, separating x from C. Complete regularity is similarly preserved. However, (T_4) is not preserved under infinite box products of spaces, as counterexamples exist where disjoint closed sets cannot be separated by disjoint opens despite normality in each factor. Regarding connectedness, finite box products of connected spaces are connected, mirroring the product topology behavior since the topologies coincide for finite indices. Path-connectedness preserves analogously under finite products. In contrast, infinite box products of connected Hausdorff spaces with infinitely many nondegenerate factors are disconnected; for instance, \mathbb{R}^\mathbb{N} in the box topology admits a disconnection into the set of bounded sequences and the set of unbounded sequences, both clopen, and further decomposes into continuum many disjoint nonempty open subsets. This follows from the abundance of open sets in the box topology, allowing separations that the coarser prevents.

Continuity of Canonical Maps

In the box topology \tau_b on the Cartesian product \prod_{i \in I} X_i, where each X_i is equipped with its topology \tau_i, the canonical projection maps \pi_k: (\prod_{i \in I} X_i, \tau_b) \to (X_k, \tau_k) are continuous for every index k \in I. To verify this, consider an V_k \in \tau_k. The preimage is \pi_k^{-1}(V_k) = \prod_{i \neq k} X_i \times V_k, where each X_i (for i \neq k) is open in itself and V_k is open in X_k. This set is a basis element of the box topology, as the basis consists of arbitrary products of open sets from each factor. Thus, \pi_k^{-1}(V_k) is open in \tau_b, confirming by the definition of . The box topology ensures that the product \prod_{i \in I} X_i serves as the categorical product in the , characterized by the continuous maps \pi_i as the universal morphisms. Specifically, for any Y and continuous maps f_i: Y \to X_i for each i \in I, there exists a unique map f: Y \to \prod_{i \in I} X_i such that \pi_i \circ f = f_i for all i, and this f is continuous with respect to the box topology when the index set I is finite; for infinite I, the construction aligns with the projections' continuity but highlights distinctions in mapping properties. Additionally, the identity map \mathrm{id}: (\prod_{i \in I} X_i, \tau_p) \to (\prod_{i \in I} X_i, \tau_b), where \tau_p denotes the , is continuous, as every basis element of \tau_p (products of opens with only finitely many non-full factors) is also a basis element of \tau_b. However, the reverse identity map \mathrm{id}: (\prod_{i \in I} X_i, \tau_b) \to (\prod_{i \in I} X_i, \tau_p) is not continuous in general when I is infinite, since basis elements of \tau_b with infinitely many non-full factors need not be open in \tau_p. Detailed counterexamples appear in the section on failure of continuity.

Compactness Behavior

In the box topology, the Cartesian product \prod_{i \in I} X_i is compact if and only if each space X_i is compact and the I is finite. When I is finite, the box topology coincides with the , and follows directly from the finite-product case of , which states that the product of finitely many compact spaces is compact. For infinite index sets, however, an adaptation of fails to hold; the box topology does not preserve even when every factor is compact, distinguishing it sharply from the product topology where arbitrary products of compact spaces remain compact. Local compactness is similarly preserved under the box topology only in the finite-product case: if each X_i is locally compact and I is finite, then \prod_{i \in I} X_i is locally compact in the box topology. For infinite products, even when all factors are locally compact, the resulting box product is never locally compact, as non-trivial infinite box products lack the necessary compact neighborhoods around points. In the countable case, where I = \mathbb{N}, the box and product topologies on products of compact spaces do not coincide with respect to compactness preservation. While the product topology yields a compact space by , the box topology does not, as illustrated by the specific fact that [0,1]^\mathbb{N} fails to be compact despite each factor [0,1] being compact. This underscores the stricter open covers in the box topology, where basis elements require openness in every coordinate simultaneously, preventing finite subcovers for certain infinite collections.

Convergence Criteria

In the box topology on a product \prod_{i \in I} X_i, convergence of a (x_\lambda)_{\lambda \in \Lambda} to a point x \in \prod_{i \in I} X_i requires that for every basis element \prod_{i \in I} U_i containing x, where each U_i is open in X_i with x_i \in U_i, there exists \lambda_0 \in \Lambda such that x_\lambda \in \prod_{i \in I} U_i for all \lambda \geq \lambda_0. This is equivalent to \pi_i(x_\lambda) \in U_i for all i \in I and \lambda \geq \lambda_0. A necessary condition for such is coordinatewise : \pi_i(x_\lambda) \to \pi_i(x) in X_i for every i \in I, since the projection maps \pi_i are continuous from the product with the box topology to X_i. However, coordinatewise is not sufficient when I is , as the simultaneous entry into arbitrary box neighborhoods demands a form of uniformity across all coordinates. For filters, convergence in the box topology to x means that every box neighborhood of x belongs to the filter \mathcal{F}. This condition is equivalent to the projected filter \pi_i(\mathcal{F}) converging to \pi_i(x) in each X_i, but again, the converse requires that the filter refines the box neighborhood filter of x, which is stricter than refining the product neighborhood filter when I is infinite. The box topology thus enforces convergence in every coordinate topology while ensuring the filter adheres to neighborhoods that constrain all coordinates without finite support exceptions. Regarding sequences, which are special cases of nets indexed by \mathbb{N}, convergence in the box topology coincides with convergence in the product topology only for finite products, where the topologies agree. For countable infinite products (e.g., I = \mathbb{N}), sequential convergence in the box topology is stricter: while the product topology requires only coordinatewise convergence, the box topology demands that changes occur in only finitely many coordinates eventually. Specifically, if each X_i is Hausdorff, a sequence (x_n) converges to x if there exists a finite subset J \subset I and n_0 \in \mathbb{N} such that x_n|_J \to x|_J pointwise in the finite product topology on \prod_{j \in J} X_j, and x_n(i) = x(i) for all i \notin J and n > n_0. For uncountable I, the condition is even more restrictive, resembling uniform stabilization across coordinates, preventing sequences with infinitely supported variations from converging. This contrasts with the product topology, where finite changes suffice for tails in neighborhoods, but infinite support alterations (as in the sequence where the n-th term differs from the limit in all coordinates beyond n) block convergence in the box topology.

Illustrative Examples

Failure of Continuity

One key distinction between the product and box topologies arises in the of maps between spaces equipped with these topologies. When the index set is finite, the product topology and box topology coincide on the product space, making the identity map continuous in both directions. However, for infinite s, the box topology is strictly finer than the , and the identity map from the product topology to the box topology fails to be continuous. A classic example illustrates this discontinuity on \mathbb{R}^\mathbb{N}. Consider the set U = \prod_{n=1}^\infty \left(-\frac{1}{n}, \frac{1}{n}\right), which is open in the box topology as it is a basic open set with nonempty open intervals in each coordinate. The preimage of U under the identity map \mathrm{id}: (\mathbb{R}^\mathbb{N}, \text{[product topology](/page/Product_topology)}) \to (\mathbb{R}^\mathbb{N}, \text{box topology}) is U itself. However, U is not open in the product topology because every basic open set in the product topology intersects only finitely many coordinates nontrivially (with the rest being all of \mathbb{R}), and no such finite intersection can be contained within U, as points in U require simultaneous boundedness in all coordinates by shrinking intervals. Thus, \mathrm{id}^{-1}(U) is not open in the product topology, so the identity map is discontinuous. This failure stems from the differing neighborhood structures: basic open sets in the box topology demand uniform control across all coordinates simultaneously, whereas those in the product topology allow "finite support" with arbitrary behavior in all but finitely many coordinates. Another example is the diagonal map D: [0,1] \to [0,1]^\mathbb{N}, defined by D(t) = (t, t, t, \dots), which is continuous when [0,1]^\mathbb{N} has the (as each coordinate projection D_i(t) = t is continuous) but discontinuous in the box topology. For the open set V = \prod_{n=1}^\infty (0, 1/2) in the box topology, D^{-1}(V) = (0, 1/2), which is open in [0,1], but more refined box neighborhoods around D(0) or boundary points reveal the lack of matching preimages due to the infinite simultaneous constraints.

Failure of Compactness

A classic illustration of the failure of compactness in the box topology arises with the product space [0,1]^I, where I is an uncountable index set such as the real numbers \mathbb{R}. By Tychonoff's theorem, this space is compact under the product topology. However, the box topology renders it non-compact. This non-compactness can be demonstrated by considering the subspace S = \{0,1\}^I \subseteq [0,1]^I, consisting of all characteristic functions from I to \{0,1\}. The subspace topology on S induced by the box topology on [0,1]^I coincides with the box topology on the product \{0,1\}^I, where \{0,1\} carries the discrete topology. In this topology, every singleton is open, as it is the basic open set \prod_{i \in I} \{f(i)\} for each function f: I \to \{0,1\}. Thus, S is an uncountable discrete space. The collection of all singletons forms an open cover of S with no finite subcover, so S is not compact. Since a non-compact subspace implies the ambient space [0,1]^I is not compact, the box topology fails to preserve compactness here. A sketch of the general reason for this failure is that compact subsets in the box topology require simultaneous boundedness across all coordinates: for any compact K, there must exist uniform bounds on the "oscillation" or extent in every direction, meaning K lies in a product of sets with diameters controlled independently yet uniformly for the entire set. This condition holds for finite I but fails for infinite I, as points in the full product vary freely in infinitely many coordinates without uniform restraint, preventing finite subcovers for covers that probe all coordinates equally. In the countable case with I = \mathbb{N}, the box topology on [0,1]^\mathbb{N} also fails for the same reason: \{0,1\}^\mathbb{N} is uncountable and in the induced box topology. The product and box topologies coincide only for finite I, in which case holds in both; for infinite I, the box topology is rarely compact on such products.

Relation to Product Topology

Inclusion and Coarseness

The box topology on a product space \prod_{i \in I} X_i is finer than the , meaning that the collection of open sets in the is a of those in the box topology, denoted \tau_p \subseteq \tau_b. This arises because the basis for the consists of sets of the form \prod_{i \in I} U_i, where each U_i is open in X_i and U_i = X_i for all but finitely many i \in I; such sets are also basis elements for the box topology, which allows U_i \neq X_i for every i \in I. Consequently, every open set in the can be expressed as a of box basis elements, but the converse does not hold in general. When the index set I is infinite, the inclusion is strict: the box topology contains open sets that are not open in the product topology. For example, consider the countable infinite product \mathbb{R}^\mathbb{N}, where each factor is the real line with its standard topology. The set U = \prod_{n=1}^\infty \left( -\frac{1}{n}, \frac{1}{n} \right) is open in the box topology because it is a product of open intervals in each \mathbb{R}. However, it is not open in the product topology, as infinitely many factors differ from \mathbb{R}, violating the finite support condition of the product basis. This demonstrates that the box topology has strictly more open sets when I is infinite. The two topologies coincide if and only if I is finite. In this case, the finite support requirement is automatically satisfied for any product of open sets, so the bases are identical and generate the same . For infinite I, the box basis properly contains the product basis, ensuring the topologies differ.

Conditions for Equality

The box \tau_b on the product space \prod_{i \in I} X_i coincides with the \tau_p if and only if the I is finite; for infinite I, the box topology is strictly finer than the . When I is finite, the basis for the product topology consists of sets of the form \prod_{i \in I} U_i, where each U_i is open in X_i and U_i = X_i for all but finitely many i; however, since I itself is finite, this condition holds vacuously for all such products, making the bases identical and thus the topologies equal. To see this explicitly, any basis element in the box topology is an arbitrary product \prod_{i \in I} U_i with each U_i open, which matches precisely the form required for the basis under the finite index restriction. In special cases with infinite I, the topologies may coincide if all but finitely many X_i carry the indiscrete topology (with only the and X_i as open sets), as the basis then reduces to products effectively determined by the finitely many non-trivial factors, aligning with the product basis; however, this equality does not hold generally for infinite products of spaces with non-trivial topologies. More broadly, in the , the finite product construction aligns the two topologies, reflecting the categorical finite products; for infinite products, the box topology imposes a "" by requiring openness in every coordinate simultaneously, diverging from the coordinatewise continuity emphasized in the .

Functional Analysis Contexts

In the framework of spaces, the box uniformity on a product \prod_{\alpha \in A} X_{\alpha} of spaces (X_{\alpha}, \mathcal{D}_{\alpha}) is generated by the base consisting of entourages \prod_{\alpha \in A} D_{\alpha}, where each D_{\alpha} \in \mathcal{D}_{\alpha}. This induces the box topology on the product and represents the finest uniformity compatible with it, ensuring that every \pi_{\alpha}: \prod_{\beta \in A} X_{\beta} \to X_{\alpha} is . Such a uniformity facilitates the of for functions defined on infinite products, as it imposes simultaneous control across all coordinates without restricting to finite subsets, unlike the coarser product uniformity. For function spaces, consider the box topology on C(X)^I, the product of copies of the space C(X) of continuous real-valued functions on a X, indexed by a set I. In this , convergence of nets corresponds to entry into arbitrary products of neighborhoods in each component C(X), which strengthens (as in the ) to require uniformity across the entire I with potentially varying neighborhood sizes per coordinate. However, this leads to a failure of joint for the map \mathrm{ev}: C(X)^I \times X \to \mathbb{R}^I, defined by (f, x) \mapsto (f_{\alpha}(x))_{\alpha \in I}, since neighborhoods in the box topology demand control over all indices simultaneously, disrupting the compact-induced uniformity needed for joint behavior. In contrast, the on C(X), when extended to products, preserves this joint by restricting to compact subsets of X, making it more suitable for applications requiring maps to be continuous in both arguments. In infinite-dimensional analysis, the box topology provides a model for "uniform" structures on spaces such as \ell^{\infty}(I), the space of bounded real sequences indexed by I, which embeds into \mathbb{R}^I equipped with the box topology to study bounded linear operators and their continuity properties under strong convergence criteria. This approach aids in examining operator algebras and dual spaces, where the box-induced uniformity highlights behaviors not captured by weaker topologies like the norm or product topologies on \ell^{\infty}(I). The box topology was introduced in foundational 1950s topology texts, such as John L. Kelley's General Topology (1955), where it filled gaps in understanding infinite products, and has since been employed in counterexamples illustrating limitations of paracompactness in box products of paracompact spaces, such as the failure of countable box products of compact metric spaces to be paracompact under certain set-theoretic assumptions. Notably, the normality of the countable box product of copies of the real line remains an open problem in set-theoretic topology, highlighting unresolved questions about separation properties in box products. A specific property in this context is that the box topology preserves metrizability for products of metric spaces only when the product is finite; for infinite index sets, the resulting space fails to be first countable, as the local character at any point equals the cardinality of the index set, precluding a countable local basis essential for metrizability.

References

  1. [1]
    [PDF] Section 19. The Product Topology
    Dec 6, 2016 · In this section we consider arbitrary products of topological spaces and give two topologies on these spaces, the box topology and the product.
  2. [2]
    [PDF] Function Spaces
    A function space is a topological space whose points are functions. There are many different kinds of function spaces, and there are usually several ...
  3. [3]
    box topology in nLab
    May 4, 2017 · The box topology is strictly finer that the Tychonoff topology. Beware that it is the Tychonoff toopology which yields the actual Cartesian product.
  4. [4]
    [PDF] Topology (H) Lecture 5
    Mar 22, 2021 · Bases and sub-bases of a topology. ¶ Topology defined via a basis. If you stare at the definitions of the topologies Tmetric, TSorgenfrey, ...
  5. [5]
    Which topological properties are preserved under taking box ...
    Jun 19, 2015 · Preserved under box products: Separation axioms such as T0,T1, Hausdorffness, regularity, complete regularity; zero-dimensionality, total ...Where else do the (topology) separation axioms turn up?Connected box products of Hausdorff spaces - MathOverflowMore results from mathoverflow.net
  6. [6]
    Is the box topology good for anything? - Math Stack Exchange
    May 2, 2011 · ... box product of normal spaces need not be normal. From a quick look through "Counterexamples in Topology" and a quick google search, problems ...Normal spaces in box and uniform topology - Math Stack ExchangeIf the product of spaces is normal, then factor are normal. [duplicate]More results from math.stackexchange.com
  7. [7]
    The (dis)connectedness of products of Hausdorff spaces in the box ...
    In this paper the following two propositions are proved: (a) If , , is an infinite system of connected spaces such that infinitely many of them are ...
  8. [8]
    Connectedness in box topology - Math Stack Exchange
    Sep 27, 2021 · For the box topology there is a sort of anti-connectedness theorem: any infinite family of Xi,i∈I where all Xi are T3 and have at least two ...connected components in box topology - Math Stack ExchangeBox topology and connectedness property - Math Stack ExchangeMore results from math.stackexchange.com
  9. [9]
    [PDF] §19 Product Topology (general case)
    -1IUΜM : ΜخL, UΜ ج XΜ open=, then the topology generated by it, is the coarsest topology containing subbasis S. ... ¥ H-1, 1L is open in the box topology, but not ...<|control11|><|separator|>
  10. [10]
    [PDF] The Product Topology—Proofs of Theorems
    Jun 12, 2016 · Theorem 19.1. Comparison of the Box and Product Topologies. The box topology on nα∈J Xα has as a basis all sets of the form nα∈J Uα.
  11. [11]
    [PDF] 8. Finite Products
    Therefore any topology in which the projections are continuous ... This second definition is called the product topology, while the first one (generated by the ...
  12. [12]
    245B, Notes 10: Compactness in topological spaces - Terry Tao
    Feb 9, 2009 · The box topology usually does not preserve compactness. For instance, one easily checks that the product of any number of discrete spaces is ...
  13. [13]
    [0,1]N with respect to the box topology is not compact
    Aug 28, 2011 · A topological space X is said to be countably compact if every countable open cover of X has a finite subcover. Is [0,1]N countably compact in the box topology?Compact topology between box and product topologybox topology vs product topology - Math Stack ExchangeMore results from math.stackexchange.com
  14. [14]
    [PDF] Convergence in the box topology (Problem Set 2, #5)
    So the box topology isn't just strong—it is so strong that it prevents almost all reasonable sequences from converging. In that sense, speaking informally, it ...
  15. [15]
    [PDF] Assignment #2 Solutions
    Jun 9, 2023 · Since the product topology is coarser than the box topology, convergence in the product topology is weaker. (As you can notice from the ...
  16. [16]
    [PDF] Chapter 9 Convergence
    Both nets and filters are sufficient to describe the topology in any space, so we should be able to use them to describe continuous functions. Theorem 5.12 ...
  17. [17]
    2.2 Product Spaces - Math 581: Topology 1
    Remark. The box topology on X is finer than the product topology and when A is finite, these topologies are identical.
  18. [18]
    [PDF] 17. Infinite products Definition-Lemma 17.1. Let Xα, α ∈ Λ be a ...
    We put a topology on X by taking the smallest topology such all of these projection maps are continuous. ... to X with the box topology is not continuous.
  19. [19]
    [PDF] Topology - Harvard Mathematics Department
    3 Topology. This section describes the basic definitions and constructions in topology, together with lots of examples. Topology. A topological space is a ...
  20. [20]
    [PDF] CLASS NOTES MATH 551 (FALL 2014) 1. Wed, Aug. 27 Topology is ...
    (1) (Covering property) Every point of x lies in at least one basis element (2) (Intersection property) If B1,B2 ∈ B and x ∈ B1 ∩ B2, then there is a third ...
  21. [21]
    [PDF] TYCHONOFF THEOREM AND ITS APPLICATIONS Last time
    Mar 7, 2021 · We have seen the failure for sequential compactness. (3) According to Tychonoff theorem, ([0,1][0,1],Tproduct) is compact. Thus. Corollary ...
  22. [22]
    [PDF] GeneralTopologyWillard.pdf - Rexresearch1.com
    ... box topology, so in those cases where we have any intuition to begin with ... net convergence is modeled after the definition of sequential convergence ...
  23. [23]
    [PDF] MTH 304: General Topology
    IV.Separation Axioms. 62. 1. Regular Spaces ... connected with the box topology, so cannot be path connected. (See Example ...
  24. [24]
    [PDF] Lecture 1: August 23 Introduction. Topology grew out of certain ...
    Aug 23, 2025 · The box topology on X is the topology generated by the basis. Y i∈I. Ui. Ui ∈ Ti for every i ∈ I . It is not hard to see that this is indeed ...<|control11|><|separator|>
  25. [25]
    [PDF] Notes on Introductory Point-Set Topology - Cornell Mathematics
    There is a more refined way to apply this fact that connectedness is preserved by homeomorphisms, using the following idea. In a connected space X , a point x ∈ ...
  26. [26]
    [PDF] Uniform box powers and products
    For instance, the unique uniformity on the compact metrizable space ω + 1 can serve as a pivot for uniform spaces with countable bases, and thus for all uniform ...
  27. [27]
    [PDF] Topology Proceedings 15 (1990) pp. 135-141
    o. INTRODUCTION. One of the outstanding open questions in set theoretic topol ogy is the question of the paracompactness of box products. The box product of ...