This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Digital topological groups

Dae-Woong Lee and P. Christopher Staecker Department of Mathematics, and Institute of Pure and Applied Mathematics, Jeonbuk National University, 567 Baekje-daero, Deokjin-gu, Jeonju-si, Jeollabuk-do 54896, Republic of Korea [email protected] Mathematics Department, Fairfield University, 1703 North Benson Rd, Fairfield, CT 06824-5195 [email protected]
Abstract.

In this article, we develop the basic theory of digital topological groups. The basic definitions directly lead to two separate categories, based on the details of the continuity required of the group multiplication. We define NP1\textrm{NP}_{1}- and NP2\textrm{NP}_{2}-digital topological groups, and investigate their properties and algebraic structure. The NP2\textrm{NP}_{2} category is very restrictive, and we give a complete classification of NP2\textrm{NP}_{2}-digital topological groups. We also give many examples of NP1\textrm{NP}_{1}-digital topological groups. We define digital topological group homomorphisms, and describe the digital counterpart of the first isomorphism theorem.

Key words and phrases:
digital image; normal product adjacency; NPi\textrm{NP}_{i}-digital topological group; digital simple closed curve; regular graph; vertex-transitive graph; Cayley graph; digital topological group homomorphism; digital open map; digital H-space
2020 Mathematics Subject Classification:
Primary 22A30; Secondary 22A10; 68U03; 05C10
The first author was supported by the National Research Foundation of Korea (NRF) grant funded by the Korean government (MSIT) (No. 2018R1A2B6004407)

1. Introduction

The topological theory of digital images focuses on topological properties of finite subsets of the nn-dimensional integer lattice, attempting to build a digital theory which resembles classical topology of n\mathbb{R}^{n}.

A graph-theoretical approach to digital topology was introduced by A. Rosenfeld in the 1960s. Interesting works on digital surfaces and digital manifolds as a special kind of combinatorial manifolds have been produced in the 1980s and in 1990s by many authors.

A classical topological group is both a topological space and a group in which the topological structure is compatible with the group structure; that is, the product and inverse are topologically continuous. A topological group is both a special case of an H-space and a general case of a Lie group in pure mathematics.

A theory of H-spaces has been developed in the digital setting by O. Ege and I. Karaca [9] and the first author in [11, 12]. The additional algebraic structure of groups allows the present paper to prove more specific results. This paper is organized as follows: In Section 2, we review and describe fundamental properties of digital images with some adjacency relations, digital continuous functions, and some terminology from graph theory. We define NPi\textrm{NP}_{i}-digital topological groups for some i{1,2}i\in\{1,2\} and digital simple closed curves. In Section 3, we investigate properties of digital topological groups, and construct many classes of examples. In Section 4, we develop digital topological group homomorphisms as morphisms of the category of digital topological groups. We also describe the digital counterpart of the first isomorphism theorem from group theory. In Section 5, we describe a complete classification of NP2\textrm{NP}_{2}-digital topological groups.

2. Preliminaries

A digital image (X,κ)(X,\kappa) consists of a finite set XX of points in n\mathbb{Z}^{n} with some adjacency relation κ\kappa which is both antireflexive and symmetric. Typically the adjacency relation is based on some notion of adjacency of points in n\mathbb{Z}^{n}. This style of digital topology has its origins in the work of Rosenfeld and others, see [15] for an early work.

When n2n\geq 2 there is no canonical “standard adjacency” to use in n\mathbb{Z}^{n} which corresponds naturally to the standard topology of n\mathbb{R}^{n}. In the case of 2\mathbb{Z}^{2}, for example, at least two different adjacency relations are natural: we can view 2\mathbb{Z}^{2} as a rectangular lattice connected by the coordinate grid, so that each point is adjacent to 4 neighbors; or we can additionally allow diagonal adjacencies so that each point is adjacent to 8 neighbors.

In n\mathbb{Z}^{n}, there are nn different standard adjacency relations, denoted cuc_{u} for u{1,2,,n}u\in\{1,2,\dots,n\}, defined as follows: Two distinct points x=(x1,x2,,xn)x=(x_{1},x_{2},\ldots,x_{n}) and y=(y1,y2,,yn)y=(y_{1},y_{2},\ldots,y_{n}) in n\mathbb{Z}^{n} are cuc_{u}-adjacent [5] if

  • there are at most uu distinct indices ii such that |xiyi|=1|x_{i}-y_{i}|=1; and

  • for all other indices jj, if |xjyj|1|x_{j}-y_{j}|\neq 1, then xj=yjx_{j}=y_{j}.

We will make use of the notation xκyx\leftrightarrow_{\kappa}y when xx is adjacent to yy by the adjacency relation κ\kappa in n\mathbb{Z}^{n}, and xκyx\leftrightarroweq_{\kappa}y when xx is adjacent or equal to yy. The particular adjacency relation will usually be clear from context, and in this case we will omit the subscript.

Let aa and bb be nonnegative integers with a<ba<b. A digital interval [2] is a finite set of the form

[a,b]={z|azb},[a,b]_{\mathbb{Z}}=\{z\in\mathbb{Z}~{}|~{}a\leq z\leq b\},

with the c1c_{1}-adjacency relation in \mathbb{Z}.

A digital image (X,κ)(X,\kappa) is said to be κ\kappa-connected [15] if for every pair x,yx,y of points of XX, there exists a subset {x0,x1,,xs}X\{x_{0},x_{1},\ldots,x_{s}\}\subseteq X

  • x=x0x=x_{0};

  • xs=yx_{s}=y; and

  • xixi+1x_{i}\leftrightarroweq x_{i+1} for i=0,1,2,,s1i=0,1,2,\ldots,s-1 for s1s\geq 1.

Such a set {x0,x1,,xs}\{x_{0},x_{1},\ldots,x_{s}\} is called a path (or κ\kappa-path) from xx to yy. Given some xXx\in X, the set of all points of XX having a path to xx is the connected component of xx. A digital image is digitally connected if and only if it consists of a single component.

A function f:(X,κ)(Y,λ)f:(X,\kappa)\rightarrow(Y,\lambda) between digital images is called a (κ,λ)(\kappa,\lambda)-continuous function [3] when for any x1,x2Xx_{1},x_{2}\in X, if x1x2x_{1}\leftrightarrow x_{2}, then f(x1)f(x2)f(x_{1})\leftrightarroweq f(x_{2}). Equivalently, ff is continuous if f(C)f(C) is a λ\lambda-connected subset of YY for each κ\kappa-connected subset CC of XX. When the adjacencies are clear, we simply say that ff is continuous. Any composition of digitally continuous functions is digitally continuous.

A (κ,λ)(\kappa,\lambda)-continuous function of digital images

f:(X,κ)(Y,λ)f:(X,\kappa)\rightarrow(Y,\lambda)

is called a (κ,λ)(\kappa,\lambda)-isomorphism if ff is a bijection set-theoretically, and its inverse f1:(Y,λ)(X,κ)f^{-1}:(Y,\lambda)\rightarrow(X,\kappa) is (λ,κ)(\lambda,\kappa)-continuous. In this case, (X,κ)(X,\kappa) and (Y,λ)(Y,\lambda) are said to be (κ,λ)(\kappa,\lambda)-isomorphic; see [2] and [4]. We often make use of the following fact: if ff is an isomorphism, then xyx\leftrightarrow y if and only if f(x)f(y)f(x)\leftrightarrow f(y).

Any digital image (X,κ)(X,\kappa) can naturally be viewed as a finite simple graph with vertex set XX and an edge connecting x,yXx,y\in X if and only if xyx\leftrightarrow y. Conversely, any graph may be considered as a digital image with a standard adjacency:

Theorem 2.1 ([13], Proposition 2.5).

Let XX be a finite simple graph of kk vertices. Then there is some n<kn<k such that XX may be embedded as a digital image X[1,1]nX\subset[-1,1]_{\mathbb{Z}}^{n} with cnc_{n}-adjacency.

(Above, by “embedded” we mean that XX, considered as an abstract graph, is isomorphic as a graph to the digital image X[1,1]nX\subset[-1,1]_{\mathbb{Z}}^{n} with cnc_{n}-adjacency.)

We will, whenever convenient, describe a digital image (X,κ)(X,\kappa) in graph-theoretic terms. Specifically, our definitions above of paths, connectedness, and components correspond exactly to the same concepts in graph theory. We will also make use of the degree from graph theory: for xXx\in X, the degree of xx is the number of yXy\in X with yxy\leftrightarrow x. If all points of XX have the same degree dd, we say XX is dd-regular. If all pairs of points of XX are adjacent, then XX is a complete graph.

Given two digital images X1X_{1} and X2X_{2}, we can consider the Cartesian product X1×X2X_{1}\times X_{2} as a digital image, but there are several natural choices for the adjacency relations to be used in the Cartesian product, analogous to the various cuc_{u}-adjacencies. The most natural product adjacencies are the normal product adjacencies, which were defined by Boxer as follows:

Definition 2.2.

([6]) Let {(Xi,κi)i=1,2,,n}\{(X_{i},\kappa_{i})\mid i=1,2,\cdots,n\} be an indexed family of digital images. Then for some u{1,2,,n}u\in\{1,2,\dots,n\}, the normal product adjacency NPu(κ1,κ2,,κn)\textrm{NP}_{u}(\kappa_{1},\kappa_{2},\dots,\kappa_{n}) is the adjacency relation on i=1nXi\prod_{i=1}^{n}X_{i} defined by: (x1,x2,,xn)(x_{1},x_{2},\dots,x_{n}) and (x1,x2,,xn)(x^{\prime}_{1},x^{\prime}_{2},\dots,x^{\prime}_{n}) are adjacent if and only if their coordinates are adjacent in at most uu positions, and equal in all other positions.

In this paper, our products generally have the form X×XX\times X for some digital image X:=(X,κ)X:=(X,\kappa). On such a Cartesian product, the two natural adjacencies to choose from are NP1(κ,κ)\textrm{NP}_{1}(\kappa,\kappa) and NP2(κ,κ)\textrm{NP}_{2}(\kappa,\kappa). When unambiguous, we will abbreviate NPi(κ,κ)\textrm{NP}_{i}(\kappa,\kappa) as simply NPi\textrm{NP}_{i} for i{1,2}i\in\{1,2\}. Clearly, if two points (x1,x2)(x_{1},x_{2}) and (y1,y2)(y_{1},y_{2}) in X×XX\times X are NP1\textrm{NP}_{1}-adjacent, then they are also NP2\textrm{NP}_{2}-adjacent. As we will see, the topological structure of the Cartesian product depends strongly on the choice between NP1\textrm{NP}_{1} and NP2\textrm{NP}_{2}; see [13, 17] for a discussion of NP2\textrm{NP}_{2}-homotopy relations.


3. Digital topological groups

3.1. Definitions and basic examples

A classical topological group is both a topological space and a group whose binary operation and inverse are topologically continuous. The following is the digital counterparts of classical topological groups as the topological structure of the Cartesian product depending strongly on the choice between NP1\textrm{NP}_{1} and NP2\textrm{NP}_{2}.

Definition 3.1.

Let (G,κ)(G,\kappa) be a digital image such that GG is a group with multiplication μG:G×GG\mu_{G}:G\times G\to G and inverse ιG:GG\iota_{G}:G\to G. For some i{1,2}i\in\{1,2\}, we say GG is an NPi\textrm{NP}_{i}-digital topological group when ιG\iota_{G} is (κ,κ)(\kappa,\kappa)-continuous and μG\mu_{G} is (NPi(κ,κ),κ)(\textrm{NP}_{i}(\kappa,\kappa),\kappa)-continuous. If GG is an NPi\textrm{NP}_{i}-digital topological group for some i{1,2}i\in\{1,2\}, we say GG is a digital topological group.

For a digital topological group GG with product μG\mu_{G} and inverse ιG\iota_{G} we typically will write the product of elements xx and yy as μG(x,y)=xy\mu_{G}(x,y)=xy, and the inverse of an element xx as ιG(x)=x1\iota_{G}(x)=x^{-1}.

Clearly any NP2\textrm{NP}_{2}-digital topological group is also an NP1\textrm{NP}_{1}-digital topological group. In classical topology, the NP2\textrm{NP}_{2} continuity condition is analogous to requiring that μG\mu_{G} be continuous as a function of 2 variables, while the NP1\textrm{NP}_{1} condition is analogous to requiring that μG\mu_{G} is continuous in each variable separately. In classical topology the definition of topological group requires full continuity, while a structure with continuity in each variable separately is called a semitopological group [10, page 27].

Classically, the assumption of full continuity is generally more interesting, but we will see in our setting that the more interesting examples of digital topological groups are NP1\textrm{NP}_{1} but not NP2\textrm{NP}_{2}. The full NP2\textrm{NP}_{2} continuity assumption is too strong to allow for any nontrivial examples, as we show in Section 5.

The simplest nontrivial example of an NP1\textrm{NP}_{1}-digital topological group is a simple closed curve.

Definition 3.2.

A digital simple closed curve is a set X:={x0,x1,,xn1}X:=\{x_{0},x_{1},\dots,x_{n-1}\} of distinct elements for n3n\geq 3 with some adjacency such that xixjx_{i}\leftrightarrow x_{j} if and only if j=i±1j=i\pm 1, with subscripts read modulo nn.

When viewed as a graph, all points of a digital simple closed curve have degree 2; that is, it is a connected 22-regular graph.

For example, the set [1,1]2{(0,0)}[-1,1]^{2}_{\mathbb{Z}}-\{(0,0)\} forms a digital simple closed curve in (2,c1)(\mathbb{Z}^{2},c_{1}). It is not a digital simple closed curve in (2,c2)(\mathbb{Z}^{2},c_{2}) because there are some extra diagonal adjacencies. Specifically the point (1,0)(1,0) is c2c_{2}-adjacent to its two c1c_{1}-neighbors (1,1)(1,1) and (1,1)(1,-1), but is also adjacent to (0,1)(0,1) and (0,1)(0,-1).

Now we show that a simple closed curve is an NP1\textrm{NP}_{1}-digital topological group, but not an NP2\textrm{NP}_{2}-digital topological group.

Example 3.3.

Let G={x0,x1,,xn1}G=\{x_{0},x_{1},\dots,x_{n-1}\} be a digital simple closed curve with adjacency κ\kappa, and define operations:

μG(xi,xj)=xi+j,ιG(xi)=xi,\mu_{G}(x_{i},x_{j})=x_{i+j},\qquad\iota_{G}(x_{i})=x_{-i},

where all subscripts are read modulo nn.

We will show that these operations make GG into an NP1\textrm{NP}_{1}-digital topological group. The group axioms are clearly satisfied by these operations, so we need only check that μG\mu_{G} is (NP1,κ)(\textrm{NP}_{1},\kappa)-continuous, and ιG\iota_{G} is (κ,κ)(\kappa,\kappa)-continuous.

For ιG\iota_{G}, take two κ\kappa-adjacent points xixi+1x_{i}\leftrightarrow x_{i+1}. Then

ιG(xi)=xixi1=ιG(xi+1)\iota_{G}(x_{i})=x_{-i}\leftrightarrow x_{-i-1}=\iota_{G}(x_{i+1})

as desired.

For μG\mu_{G}, we need to choose any two pairs of points in G×GG\times G which are NP1\textrm{NP}_{1}-adjacent, and show that they map by μG\mu_{G} into κ\kappa-adjacent points of GG. Points which are NP1\textrm{NP}_{1}-adjacent are equal in one coordinate, and adjacent in the other. So without loss of generality we will consider the pair (xi,xj)(xi+1,xj)(x_{i},x_{j})\leftrightarrow(x_{i+1},x_{j}). Then we have:

μG(xi,xj)=xi+jxi+j+1=μG(xi+1,xj)\mu_{G}(x_{i},x_{j})=x_{i+j}\leftrightarrow x_{i+j+1}=\mu_{G}(x_{i+1},x_{j})

as desired. We have shown that GG is an NP1\textrm{NP}_{1}-digital topological group.

In contrast, these operations do not make GG into an NP2\textrm{NP}_{2}-digital topological group when n>3n>3. To see this, note that (x0,x0)(x_{0},x_{0}) and (x1,x1)(x_{1},x_{1}) are NP2\textrm{NP}_{2}-adjacent, but their products μG(x0,x0)=x0\mu_{G}(x_{0},x_{0})=x_{0} and μG(x1,x1)=x2\mu_{G}(x_{1},x_{1})=x_{2} are not adjacent.

3.2. Properties of digital topological groups

In this subsection, we will investigate some interesting properties of digital topological groups. Most of the results in this section are inspired by corresponding statements in the classical theory.

Theorem 3.4.

Let G=(G,κ)G=(G,\kappa) be a digital topological group. Then, for each xGx\in G, the maps μx,νx:GG\mu_{x},\nu_{x}:G\to G given by μx(y)=xy\mu_{x}(y)=xy and νx(y)=yx\nu_{x}(y)=yx, respectively, are (κ,κ)(\kappa,\kappa)-isomorphisms.

Proof.

Since the argument for νx\nu_{x} is similar, we will prove the statement for μx\mu_{x}. Assume that GG is an NPi\textrm{NP}_{i}-digital topological group for some i{1,2}i\in\{1,2\}.

First we show that μx\mu_{x} is continuous for each xGx\in G. Let fx:GG×Gf_{x}:G\to G\times G be defined by fx(y)=(x,y)f_{x}(y)=(x,y) for each xGx\in G. Note that μx=μGfx\mu_{x}=\mu_{G}\circ f_{x}. Since fxf_{x} is (κ,NPj(κ,κ))(\kappa,\textrm{NP}_{j}(\kappa,\kappa))-continuous for both j=1j=1 and j=2j=2, and μG\mu_{G} is (NPi(κ,κ),κ)(\textrm{NP}_{i}(\kappa,\kappa),\kappa)-continuous, this means that μx\mu_{x} will be (κ,κ)(\kappa,\kappa)-continuous as desired.

By the same reason, μx1\mu_{x^{-1}} is continuous. But μx1\mu_{x^{-1}} is the inverse of μx\mu_{x}, and so μx\mu_{x} is continuous with continuous inverse as desired. ∎

We can use the above to demonstrate that, in Definition 3.1, continuity of the inverse follows automatically from continuity of the product. (This is not generally true in the classical theory, in which it is possible to have a topological space with continuous group product but discontinuous inverse.)

Lemma 3.5.

Let (G,κ)(G,\kappa) be a digital image such that GG is a group with multiplication μG:G×GG\mu_{G}:G\times G\to G and inverse ιG:GG\iota_{G}:G\to G. For some i{1,2}i\in\{1,2\}, assume that μG\mu_{G} is NPi\textrm{NP}_{i}-continuous for some i{1,2}i\in\{1,2\}. Then GG is an NPi\textrm{NP}_{i}-topological group.

Proof.

It suffices to show that the inverse is continuous: let xyx\leftrightarrow y, and we will show that x1y1x^{-1}\leftrightarrow y^{-1}.

Note that in Theorem 3.4, no reference was made to continuity of ιG\iota_{G}. Thus we may use Theorem 3.4 even though we have not assumed continuity of the inverse.

Since μx1\mu_{x^{-1}} is an isomorphism by Theorem 3.4, we may apply μx1\mu_{x^{-1}} to xyx\leftrightarrow y to obtain ex1ye\leftrightarrow x^{-1}y, where eGe\in G is the identity element. Then we apply νy1\nu_{y^{-1}} and obtain y1x1y^{-1}\leftrightarrow x^{-1} as desired. ∎

We also have a partial converse to Theorem 3.4: if μx\mu_{x} and νx\nu_{x} are isomorphisms for every xx, then GG is an NP1\textrm{NP}_{1}-digital topological group.

Theorem 3.6.

Let (G,κ)(G,\kappa) be a digital image such that GG is a group with multiplication μG:G×GG\mu_{G}:G\times G\to G and inverse ιG:GG\iota_{G}:G\to G. Assume that for each xGx\in G, the functions μx,νx:GG\mu_{x},\nu_{x}:G\to G are digital isomorphisms. Then GG is an NP1\textrm{NP}_{1}-digital topological group.

Proof.

By Lemma 3.5, we need only show that μG\mu_{G} is (NP1,κ)(\textrm{NP}_{1},\kappa)-continuous. We must choose two pairs of NP1\textrm{NP}_{1}-adjacent points in GG. Without loss of generality choose pairs (x,y)NP1(x,z)(x,y)\leftrightarrow_{\textrm{NP}_{1}}(x,z), and we will show that xyκxzxy\leftrightarrow_{\kappa}xz. Since (x,y)NP1(x,z)(x,y)\leftrightarrow_{\textrm{NP}_{1}}(x,z), we will have yzy\leftrightarrow z, and applying the isomorphism μx\mu_{x} gives xyκxzxy\leftrightarrow_{\kappa}xz as desired. ∎

An easy corollary will weaken the assumption above:

Corollary 3.7.

Let (G,κ)(G,\kappa) be a digital image such that GG is a group with multiplication μG:G×GG\mu_{G}:G\times G\to G and inverse ιG:GG\iota_{G}:G\to G. Assume that for each xGx\in G, the functions μx,νx:GG\mu_{x},\nu_{x}:G\to G are digitally continuous. Then GG is an NP1\textrm{NP}_{1}-digital topological group.

Proof.

By Theorem 3.6, we need only show that μx\mu_{x} and νx\nu_{x} are isomorphisms for every xx in GG. Clearly μx\mu_{x} and μx1\mu_{x^{-1}} are inverse functions. Since we have assumed that both are continuous, μx\mu_{x} must be an isomorphism. Similarly νx\nu_{x} is an isomorphism as desired. ∎

Since μx\mu_{x} is an isomorphism which carries the identity element ee to the element xx, the structure of GG near ee must be isomorphic to the structure of GG near any other element xx.

A graph GG is vertex-transitive if and only if, for all x,yGx,y\in G, there is a graph automorphism f:GGf:G\to G with f(x)=yf(x)=y; see [18, page 14], [8, page 69], and [1, page 15].

Theorem 3.8.

Let GG be a digital topological group. Then GG is vertex-transitive.

Proof.

Given a digital topological group GG and elements x,yGx,y\in G, by Theorem 3.4 the map

μyx1:GG\mu_{yx^{-1}}:G\to G

is a graph automorphism carrying xx to yy as required. ∎

Any vertex-transitive graph is automatically regular (all vertices have the same degree), so we have:

Corollary 3.9.

Let GG be a digital topological group. Then GG is regular.

The results above are analogous to the classical statement that a topological group is a homogeneous space [14, page 39].

We may also ask wether a digital topological group must be edge-transitive, that is, there is a graph automorphism carrying any given edge onto any other given edge; see [1, page 19] for a classical simple graph. This is not necessarily true for a digital topological group, as we will show later in Example 3.12.

The fact that a digital topological group must be regular is a significant restriction on the possible structure of a digital topological group. For example, among finite subsets of (2,c1)(\mathbb{Z}^{2},c_{1}), the only possible connected digital topological group is a digital simple closed curve.

Theorem 3.10.

Let G(2,c1)G\subset(\mathbb{Z}^{2},c_{1}) be a connected digital topological group. Then GG is a digital simple closed curve, or a set of 2 or fewer points.

Proof.

By Theorem 3.8, GG must be a dd-regular graph for some dd. Since we are using c1c_{1}-adjacency, the degree dd is at most 4. We will consider each possible value for d{0,1,2,3,4}d\in\{0,1,2,3,4\}, and conclude that the only allowable situation is when GG is a simple closed curve, or a set of 2 or fewer points.

If d=0d=0 then GG is a single point. If d=1d=1, then GG is a set of two adjacent points. For d=2d=2, any connected 22-regular graph is a simple cycle, which in our terminology is a simple closed curve. For d=4d=4, the only 44-regular subset of (2,c1)(\mathbb{Z}^{2},c_{1}) is all of 2\mathbb{Z}^{2}, and since GG is finite it cannot be 44-regular.

It remains to show that the case d=3d=3 is also impossible. To obtain a contradiction, assume that G(2,c1)G\subset(\mathbb{Z}^{2},c_{1}) is a finite 3-regular graph. Since GG is finite we may compose with a translation to assume without loss of generality that (0,0)G(0,0)\in G, and that there are no points (x,y)G(x,y)\in G with x<0x<0.

Since (0,0)G(0,0)\in G but no point of GG has negative first coordinate, we have (1,0)G(-1,0)\not\in G and thus since (0,0)(0,0) is degree 3, the neighbors of (0,0)(0,0) must be exactly {(0,1),(1,0),(0,1)}\{(0,1),(1,0),(0,-1)\}. In particular, we have (0,1)G(0,1)\in G. Inductively applying the same reasoning to (0,1)G(0,1)\in G, we see that in fact (0,y)G(0,y)\in G for every y>0y>0, and so GG is infinite which is a contradiction. ∎

3.3. Examples of digital topological groups

In this subsection, we present various example constructions of digital topological groups.

Given any abstract group GG, there are two trivial ways to realize GG as an NP2\textrm{NP}_{2}-digital topological group. First as a discrete digital image: we may embed GnG\subset\mathbb{Z}^{n} as a set having no adjacent points. Then the required continuity conditions for the product and inverse are satisfied automatically, because there are no adjacencies in the domain.

Second we can form an indiscrete digital image by embedding GnG\subset\mathbb{Z}^{n} as a complete graph (by Theorem 2.1 this is possible for nn large enough). In this case the required continuity conditions are satisfied automatically because all points of the codomain are adjacent.

Generally we will be interested in examples other than these trivial cases of discrete and indiscrete digital images.

First we show that a direct product of digital topological groups is a digital topological group:

Theorem 3.11.

For some i{1,2}i\in\{1,2\}, let (G,κ)(G,\kappa) and (H,λ)(H,\lambda) be NPi\textrm{NP}_{i}-digital topological groups. Then (G×H,NPi(κ,λ))(G\times H,\textrm{NP}_{i}(\kappa,\lambda)) is an NPi\textrm{NP}_{i}-digital topological group.

Proof.

Let μG:G×GG\mu_{G}:G\times G\to G and μH:H×HH\mu_{H}:H\times H\to H be the multiplication operations for GG and HH, respectively, and let ιG:GG\iota_{G}:G\to G and ιH:HH\iota_{H}:H\to H be the inverses for GG and HH, respectively. Then we define the product and inverse

μG×H:(G×H)×(G×H)G×H\mu_{G\times H}:(G\times H)\times(G\times H)\to G\times H

and

ιG×H:G×HG×H\iota_{G\times H}:G\times H\to G\times H

by

μG×H((g,h),(g,h))=(gg,hh),\mu_{G\times H}((g,h),(g^{\prime},h^{\prime}))=(gg^{\prime},hh^{\prime}),

and

ιG×H(g,h)=(g1,h1).\iota_{G\times H}(g,h)=(g^{-1},h^{-1}).

These operations clearly obey the group axioms, so we need only demonstrate the appropriate continuities, which unfortunately are notationally cumbersome. We will write

γ=NPi(κ,λ)\gamma=\textrm{NP}_{i}(\kappa,\lambda)

for the adjacency relation being used in G×HG\times H.

By Lemma 3.5, we need only demonstrate that μG×H\mu_{G\times H} is (NPi(γ,γ),γ)(\textrm{NP}_{i}(\gamma,\gamma),\gamma)-continuous. Assume that ((g,h),(g,h))((g,h),(g^{\prime},h^{\prime})) and ((x,y),(x,y))((x,y),(x^{\prime},y^{\prime})), elements of (G×H)×(G×H)(G\times H)\times(G\times H), are NPi(γ,γ)\textrm{NP}_{i}(\gamma,\gamma)-adjacent, and we must show that (gg,hh)(gg^{\prime},hh^{\prime}) and (xx,yy)(xx^{\prime},yy^{\prime}) are γ\gamma-adjacent. We will handle the i=1i=1 and i=2i=2 cases separately.

For i=1i=1, we may assume without loss of generality that (g,h)=(x,y)(g,h)=(x,y) and (g,h)γ(x,y)(g^{\prime},h^{\prime})\leftrightarrow_{\gamma}(x^{\prime},y^{\prime}). Further we can assume without loss of generality that g=xg^{\prime}=x^{\prime} and hλyh^{\prime}\leftrightarrow_{\lambda}y^{\prime}. Then when comparing (gg,hh)(gg^{\prime},hh^{\prime}) with (xx,yy)(xx^{\prime},yy^{\prime}), we see that the first coordinates are equal. In the second coordinate we have h=yh=y and hλyh^{\prime}\leftrightarrow_{\lambda}y^{\prime}, and thus hhyyhh^{\prime}\leftrightarrow yy^{\prime} since multiplication by h=yh=y is an isomorphism. Thus (gg,hh)(gg^{\prime},hh^{\prime}) and (xx,yy)(xx^{\prime},yy^{\prime}) are NP1(γ,γ)\textrm{NP}_{1}(\gamma,\gamma)-adjacent as desired.

For i=2i=2, we must additionally consider the case where ((g,h),(g,h))((g,h),(g^{\prime},h^{\prime})) and ((x,y),(x,y))((x,y),(x^{\prime},y^{\prime})) are γ\gamma-adjacent in each of the 2 coordinates. In this case we will have gκxg\leftrightarroweq_{\kappa}x, hλyh\leftrightarroweq_{\lambda}y, gκxg^{\prime}\leftrightarroweq_{\kappa}x^{\prime}, and hλyh^{\prime}\leftrightarroweq_{\lambda}y^{\prime}. By NP2(κ,κ)\textrm{NP}_{2}(\kappa,\kappa)-continuity of μG\mu_{G} and μH\mu_{H} we have ggxxgg^{\prime}\leftrightarroweq xx^{\prime} and hhyyhh^{\prime}\leftrightarroweq yy^{\prime}, and so (gg,hh)(gg^{\prime},hh^{\prime}) and (xx,hh)(xx^{\prime},hh^{\prime}) are NP2(γ,γ)\textrm{NP}_{2}(\gamma,\gamma)-adjacent (or equal) as desired. ∎

The product result above can be used to construct examples of NP1\textrm{NP}_{1}-digital topological groups other than simple closed curves. For example if XX is a simple closed curve of length 4 and YY is a simple closed curve of length 8, then X×YX\times Y is a digital topological group isomorphic to the direct product 4×8\mathbb{Z}_{4}\times\mathbb{Z}_{8}. This digital topological group is analogous to the classical torus S1×S1S^{1}\times S^{1} regarded as a topological group.

This torus serves as an example digital topological group which is not edge-transitive. Recall that a graph is edge-transitive when, given any two edges, there is a graph automorphism carrying the first to the second.

Example 3.12.

Let GG be the digital topological group which is a product of two digital simple closed curves of different lengths klk\neq l.

Viewing GG as a graph resembling the torus S1×S1S^{1}\times S^{1}, it has some edges belonging to meridians of length kk, and some belonging to meridians of length ll. Because klk\neq l, no automorphism can carry one of these edges to the other.

A large class of examples of NP1\textrm{NP}_{1}-digital topological groups is the set of Cayley graphs of groups. We will review the basic definitions as follows.

Given a finitely generated group GG, a group presentation for GG is an expression G=SRG=\langle S\mid R\rangle, where SS is a generating set for GG, and RR is a set of relations among the elements of SS. For convenience we will assume that the generating set SS is “symmetrized”, so that gSg\in S implies g1Sg^{-1}\in S.

Given such a group presentation, we form the Cayley graph [1, page 28] in which the vertex set is GG, and each element xGx\in G is connected by an edge to every other element of the form gxgx for each gSg\in S. The structure of the Cayley graph is determined by both the group GG and the set SS: the same group may give different nonisomorphic Cayley graphs depending on the choice of generating set.

Theorem 3.13.

Let XX be a Cayley graph for some group presentation G=SRG=\langle S\mid R\rangle. Then XX is an NP1\textrm{NP}_{1}-digital topological group which is isomorphic as a group to GG.

Proof.

The vertices of XX are elements of GG, so there is a natural multiplication μX:X×XX\mu_{X}:X\times X\to X and inverse ιX:XX\iota_{X}:X\to X which satisfy the group axioms. Using these operations, XX is isomorphic to GG as a group. To show that XX is an NP1\textrm{NP}_{1}-digital topological group, by Lemma 3.5 we must show that μX\mu_{X} is continuous using NP1\textrm{NP}_{1} in the domain.

Let (x,y)(x,y) and (u,v)(u,v) be NP1\textrm{NP}_{1}-adjacent, and we will show that xyuvxy\leftrightarrow uv in XX. Without loss of generality, we may assume that x=ux=u and yvy\leftrightarrow v. Since XX is a Cayley graph, yvy\leftrightarrow v means that v=gyv=gy for some gSg\in S. So we must show that xyxgyxy\leftrightarrow xgy.

By the construction of the Cayley graph we have ege\leftrightarrow g since gSg\in S. Then applying the isomorphisms μx\mu_{x} and νy\nu_{y} gives xyxgyxy\leftrightarrow xgy as desired. ∎

As a specific example, it is easy to construct the dihedral group D8D_{8} on 8 elements as an NP1\textrm{NP}_{1}-digital topological group in (3,c1)(\mathbb{Z}^{3},c_{1}) modeled on the unit cube. This is notably an example of a non-abelian digital topological group with a nontrivial (neither discrete nor indiscrete) graph structure.

eerrr2r^{2}r3r^{3}ssrsrsr2sr^{2}sr3sr^{3}s
Figure 1. A non-abelian digital topological group modeled on the dihedral group D8D_{8}.
Example 3.14.

Let G(3,c1)G\subset(\mathbb{Z}^{3},c_{1}) be the unit cube G=[0,1]3G=[0,1]_{\mathbb{Z}}^{3}. Our definition of the operations is inspired by the Cayley graph of the presentation D8=r,sr4=e,s2=e,srs1=r1D_{8}=\langle r,s\mid r^{4}=e,s^{2}=e,srs^{-1}=r^{-1}\rangle. We identify points of GG with elements of D8D_{8} according to Figure 1, and define operations

μG:G×GG\mu_{G}:G\times G\rightarrow G

and

ιG:GG\iota_{G}:G\rightarrow G

by the structure of D8D_{8}; see Tables 1 and 2.

μG\mu_{G} ee rr r2r^{2} r3r^{3} ss
ee ee rr r2r^{2} r3r^{3} ss
rr rr r2r^{2} r3r^{3} ee rsrs
r2r^{2} r2r^{2} r3r^{3} ee rr r2sr^{2}s
r3r^{3} r3r^{3} ee rr r2r^{2} r3sr^{3}s
ss ss r3sr^{3}s r2sr^{2}s rsrs ee
Table 1. The multiplication μG:G×GG\mu_{G}:G\times G\rightarrow G.
ee rr r2r^{2} r3r^{3} ss rsrs r2sr^{2}s r3sr^{3}s
ιG\iota_{G} ee r3r^{3} r2r^{2} rr ss sr3sr^{3} sr2sr^{2} srsr
Table 2. The inverse operation ιG:GG\iota_{G}:G\rightarrow G.

A typical Cayley graph will not be an NP2\textrm{NP}_{2}-digital topological group. The assumption below will typically be satisfied by any Cayley graph, as long as there is some element of order greater than 2.

Theorem 3.15.

Let XX be a Cayley graph for some group presentation G=SRG=\langle S\mid R\rangle having some generator xSx\in S such that x2Sx^{2}\not\in S and x2ex^{2}\neq e. Then XX is not an NP2\textrm{NP}_{2}-digital topological group.

Proof.

Since xSx\in S we will have exe\leftrightarrow x, and so (e,e)(e,e) and (x,x)(x,x) are NP2\textrm{NP}_{2}-adjacent. But μX(e,e)=e\mu_{X}(e,e)=e and μX(x,x)=x2\mu_{X}(x,x)=x^{2} are not equal or adjacent because x2Sx^{2}\not\in S and x2ex^{2}\neq e was assumed. Thus μX\mu_{X} is not an NP2\textrm{NP}_{2}-continuous function. ∎

It is natural to consider the converse question to Theorem 3.13: if GG is an NP1\textrm{NP}_{1}-digital topological group, then is GG a Cayley graph for some group isomorphic to GG? The answer is no, as the following example demonstrates:

Example 3.16.

Let G2G\subset\mathbb{Z}^{2} be the unit square of four points:

x0=(0,0),x1=(1,0),x2=(1,1),x3=(0,1),x_{0}=(0,0),x_{1}=(1,0),x_{2}=(1,1),x_{3}=(0,1),

and define operations

μG(xi,xj)=xi+j\mu_{G}(x_{i},x_{j})=x_{i+j}

and

ιG(xi)=xi\iota_{G}(x_{i})=x_{-i}

with subscripts read modulo 44, so that GG is isomorphic as a group to the cyclic group 4\mathbb{Z}_{4} of order 4.

The operation above makes (G,c2)(G,c_{2}) into a digital topological group, where the continuity of μG\mu_{G} and ιG\iota_{G} is automatic because (G,c2)(G,c_{2}) is a complete graph. But GG is not the Cayley graph for 4\mathbb{Z}_{4}, which would be a simple cycle graph of 4 points.

3.4. Connected components of a digital topological group

A basic result from the classical theory of topological groups is that the component GeG_{e} of the identity element ee of a topological group GG is a normal subgroup; see [14, page 39]. The same is true in the digital setting; see Theorem 3.20 below.

Definition 3.17.

Let GG be a digital topological group with identity element eGe\in G. Define GeG_{e} to be the connected component containing the identity element ee, which we call the identity component.

More generally, for some xGx\in G, let GxG_{x} be the component containing xx.

Lemma 3.18.

Let GG be a digital topological group. For x,yGx,y\in G, we have μx(Gy)=Gxy\mu_{x}(G_{y})=G_{xy} and νx(Gy)=Gyx\nu_{x}(G_{y})=G_{yx}.

Proof.

Since the statement for νx\nu_{x} is similar, we will prove the statement for μx\mu_{x}. Note that

μx(y)=xyGxy.\mu_{x}(y)=xy\in G_{xy}.

Thus μx\mu_{x} maps yy into GxyG_{xy}, and since the continuous image of a digital connected subset is connected, we must have

μx(Gy)Gxy.\mu_{x}(G_{y})\subseteq G_{xy}.

To see that these sets are equal, note that μx1\mu_{x^{-1}} maps xyxy into GyG_{y}, and so by the same reasons as above we have

μx1(Gxy)Gy.\mu_{x^{-1}}(G_{xy})\subseteq G_{y}.

Applying μx\mu_{x} gives Gxyμx(Gy)G_{xy}\subseteq\mu_{x}(G_{y}), which combines with the above to give

μx(Gy)=Gxy\mu_{x}(G_{y})=G_{xy}

as required.∎

Since μx\mu_{x} is a digital isomorphism, the above gives:

Corollary 3.19.

Let GG be a digital topological group. For x,yGx,y\in G, the sets GxG_{x} and GyG_{y} are digitally isomorphic.

The above means that any digital topological group GG naturally decomposes as a disjoint union of connected components, each isomorphic to GeG_{e}. Specifically, any component GxG_{x} is digitally isomorphic to GeG_{e} by the isomorphism μx:GeGx\mu_{x}:G_{e}\to G_{x}.

Theorem 3.20.

Let GG be a digital topological group. Then GeG_{e} is a normal subgroup of GG.

Proof.

To show that GeG_{e} is a subgroup, take x,yGex,y\in G_{e}, and we will show that x1yGex^{-1}y\in G_{e}. Since xx and yy are in the same component, we can construct a path from xx to yy in GeG_{e}. Applying the isomorphism μx1\mu_{x^{-1}} to this path gives a path from ee to x1yx^{-1}y, which means that x1yGex^{-1}y\in G_{e} as desired.

To show that GeG_{e} is normal, take xGx\in G and yGey\in G_{e}, and we must show that xyx1Gexyx^{-1}\in G_{e}. Since yGey\in G_{e}, by Lemma 3.18 we have μx(y)Gx\mu_{x}(y)\in G_{x}, and thus xyx1=νx1(μx(y))Gexyx^{-1}=\nu_{x^{-1}}(\mu_{x}(y))\in G_{e} as desired. ∎

Because GeG_{e} is a connected subgroup and the whole group GG consists of disjoint copies of GeG_{e}, our task of classifying digital topological groups will focus on connected groups.

Let CGC_{G} be the set of all (connected) components of a digital topological group GG. Suggested by Lemma 3.18, there is a natural group structure on the set CGC_{G} of components of GG as follows.

Theorem 3.21.

Let GG be a digital topological group. Then CGC_{G} is a group, with operations given by GxGy=GxyG_{x}\cdot G_{y}=G_{xy} and (Gx)1=Gx1(G_{x})^{-1}=G_{x^{-1}}, where x,yGx,y\in G.

Proof.

By the group properties of GG, these operations on CGC_{G} will satisfy the group axioms. It suffices only to show that these operations are well-defined.

To show that the inverse is well defined, take x,xx,x^{\prime} in GG with Gx=GxG_{x}=G_{x^{\prime}}, and we will show that Gx1=Gx1G_{x^{-1}}=G_{x^{\prime-1}}. Since Gx=GxG_{x}=G_{x^{\prime}}, there is a path from xx to xx^{\prime} in GG. Since the inverse map

ιG:GG\iota_{G}:G\rightarrow G

is continuous, we may apply ιG\iota_{G} to this path to obtain a path from x1x^{-1} to x1x^{\prime-1}, which shows that x1x^{-1} and x1x^{\prime-1} are in the same component, and thus that Gx1=Gx1G_{x^{-1}}=G_{x^{\prime-1}} as desired.

Now we show that the product \cdot is well-defined. To do this, take x,y,x,yGx,y,x^{\prime},y^{\prime}\in G with Gx=GxG_{x}=G_{x^{\prime}} and Gy=GyG_{y}=G_{y^{\prime}}, and we must show that Gxy=GxyG_{xy}=G_{x^{\prime}y^{\prime}}. That is, we must show that: when xx and xx^{\prime} are in the same component, and when yy and yy^{\prime} are in the same component, then xyxy and xyx^{\prime}y^{\prime} are in the same component.

Since xx and xx^{\prime} are in the same component, there is a path

x=x0,x1,,xn=xx=x_{0},x_{1},\dots,x_{n}=x^{\prime}

from xx to xx^{\prime} in GG. Similarly there is a path

y=y0,y1,,ym=yy=y_{0},y_{1},\dots,y_{m}=y^{\prime}

from yy to yy^{\prime} in GG. By perhaps repeating elements several times to lengthen the paths, we may assume that m=nm=n so that these paths have the same length. We will also assume that points are repeated in these paths in the following way:

  • for even ii we assume that xi=xi+1x_{i}=x_{i+1}, and

  • or odd ii we assume that yi=yi+1y_{i}=y_{i+1}.

Now we may multiply these paths pointwise to obtain:

xy=x0y0,x1y1,,xnyn=xyxy=x_{0}y_{0},x_{1}y_{1},\dots,x_{n}y_{n}=x^{\prime}y^{\prime}

and it suffices to show that this sequence of points forms a path from xyxy to xyx^{\prime}y^{\prime}. We must demonstrate that

xiyixi+1yi+1x_{i}y_{i}\leftrightarroweq x_{i+1}y_{i+1}

for each i<ni<n. Because of the structure of repeated values in even and odd positions, we see that (xi,yi)(x_{i},y_{i}) and (xi+1,yi+1)(x_{i+1},y_{i+1}) are NP1\textrm{NP}_{1}-adjacent for any i<ni<n. Thus their products xiyix_{i}y_{i} and xi+1yi+1x_{i+1}y_{i+1} are adjacent as desired. ∎

Theorem 3.22.

Let GG be a digital topological group with identity element ee. Then there is a group isomorphism G/GeCGG/G_{e}\cong C_{G}.

Proof.

For xGx\in G, let x¯G/Ge\bar{x}\in G/G_{e} be the representative of xx in the quotient. Let

f:G/GeCGf:G/G_{e}\to C_{G}

be given by

f(x¯)=Gx.f(\bar{x})=G_{x}.

We will show that ff is an isomorphism.

First we show that ff is well-defined. Let x¯=y¯\bar{x}=\bar{y}, and we must show that Gx=GyG_{x}=G_{y}, that is, xx and yy are in the same component. The fact that x¯=y¯\bar{x}=\bar{y} means that x=gyx=gy for some gGeg\in G_{e}. Since gGeg\in G_{e} and yGyy\in G_{y}, we have gyGygy\in G_{y} by Lemma 3.18. Thus xGyx\in G_{y}, and so xx and yy are in the same component as desired.

Clearly ff is a group homomorphism because

f(x¯y¯)=Gxy=GxGy=f(x¯)f(y¯)f(\bar{x}\bar{y})=G_{xy}=G_{x}\cdot G_{y}=f(\bar{x})f(\bar{y})

and

f(x¯1)=Gx1=(Gx)1=f(x¯)1.f(\bar{x}^{-1})=G_{x^{-1}}=(G_{x})^{-1}=f(\bar{x})^{-1}.

Finally we show that ff is a bijection. The kernel of ff is any x¯\bar{x} with Gx=GeG_{x}=G_{e}, that is to say x¯kerf\bar{x}\in\ker f if and only if xGex\in G_{e}, which holds only when xx is trivial in G/GeG/G_{e}. Thus ff has trivial kernel. Finally ff is surjective because for any element GxCGG_{x}\in C_{G}, we have

f(x¯)=Gxf(\bar{x})=G_{x}

as required. ∎

Note that Theorem 3.22 asserts that there is a short exact sequence of groups

(3)

that is not necessarily split. If the sequence (3) happens to be split, then GG is a semidirect product of CGC_{G} and GeG_{e}; that is,

GGeCG.G\cong G_{e}\rtimes C_{G}.

In particular, it is not always true that GG is algebraically the direct product GGe×CGG\cong G_{e}\times C_{G}, as the following example shows:

Example 3.23.

Let GG be the dihedral group D8D_{8} (see Example 3.14), this time modeled on the set

G=({0,1}×{0,1})({3,4}×{0,1})(2,c1)G=(\{0,1\}\times\{0,1\})\cup(\{3,4\}\times\{0,1\})\subset(\mathbb{Z}^{2},c_{1})

as shown in Figure 2.

eerrr2r^{2}r3r^{3}ssrsrsr2sr^{2}sr3sr^{3}s
Figure 2. Another realization of D8D_{8} as an NP1\textrm{NP}_{1}-digital topological group.

We must show that the product operation used in Example 3.14 is continuous with respect to the adjacencies in the figure.

An arbitrary point of GG has the form rksϵr^{k}s^{\epsilon}, where k{0,1,2,3}k\in\{0,1,2,3\} and ϵ{0,1}\epsilon\in\{0,1\}. To show that the product is continuous, we choose two pairs of NP1\textrm{NP}_{1}-adjacent points. Without loss of generality, we take the pairs (rksϵ,rlsδ)(r^{k}s^{\epsilon},r^{l}s^{\delta}) and (rksϵ,rl+1sδ)(r^{k}s^{\epsilon},r^{l+1}s^{\delta}) where k,l{0,1,2,3}k,l\in\{0,1,2,3\} and ϵ,δ{0,1}\epsilon,\delta\in\{0,1\}, where exponents of rr are always read modulo 4, and exponents of ss are always read modulo 2. We must show that the products rksϵrlsδr^{k}s^{\epsilon}r^{l}s^{\delta} and rksϵrl+1sδr^{k}s^{\epsilon}r^{l+1}s^{\delta} are adjacent in GG.

When ϵ=0\epsilon=0, we have

rksϵrlsδ=rk+lsδrk+l+1sδ=rksϵrl+1sδr^{k}s^{\epsilon}r^{l}s^{\delta}=r^{k+l}s^{\delta}\leftrightarrow r^{k+l+1}s^{\delta}=r^{k}s^{\epsilon}r^{l+1}s^{\delta}

as desired. Recall that in D8D_{8} we have srs1=r1srs^{-1}=r^{-1}, which means that sr=r1ssr=r^{-1}s and more generally that srk=rkssr^{k}=r^{-k}s. Thus when ϵ=1\epsilon=1, we have

rksϵrlsδ=rkrlssδrkrl1ssδ=rksϵrl+1sδr^{k}s^{\epsilon}r^{l}s^{\delta}=r^{k}r^{-l}ss^{\delta}\leftrightarrow r^{k}r^{-l-1}ss^{\delta}=r^{k}s^{\epsilon}r^{l+1}s^{\delta}

as desired.

Thus GG is an NP1\textrm{NP}_{1}-digital topological group. Observe that GeG_{e} is isomorphic as a group to 4\mathbb{Z}_{4}, while CGC_{G} is isomorphic to the group 2\mathbb{Z}_{2}, and so GG is not the direct product of GeG_{e} and CGC_{G}. Rather, it is the semidirect product D842D_{8}\cong\mathbb{Z}_{4}\rtimes\mathbb{Z}_{2}, where the 2\mathbb{Z}_{2} factor acts by inversion.

This example also demonstrates that digital topological groups which are isomorphic as groups may not be isomorphic as digital images. The present example and Example 3.14 are algebraically the same group, but realized as two different digital images.


4. Digital topological group homomorphisms

In this section, we consider morphisms between digital topological groups, and the first isomorphism theorem for digital topological groups. We define a morphism between digital topological groups as follows.

Definition 4.1.

Let G=(G,κ)G=(G,\kappa) and H=(H,λ)H=(H,\lambda) be digital topological groups. A map f:GHf:G\rightarrow H is called a digital topological group homomorphism if it is both a (κ,λ)(\kappa,\lambda)-continuous function and a group homomorphism.

Definition 4.2.

A digital topological group homomorphism f:GHf:G\rightarrow H is called a digital topological group isomorphism if it is a digital isomorphism; that is, there exists a digital topological group homomorphism f1:HGf^{-1}:H\rightarrow G such that

f1f=1Gf^{-1}\circ f=1_{G}

and

ff1=1H.f\circ f^{-1}=1_{H}.

In this case, GG is said to be digital topological group isomorphic to HH.

Let NN be a subgroup, group-theoretically, of a digital topological group GG. We define an adjacency relation on NN as follows.

Definition 4.3.

Let NN be a subgroup of a digital topological group G=(G,κ)G=(G,\kappa). We define an adjacency relation κ|N\kappa|_{N} on NN by the restriction of κ\kappa to NN. Using this restricted adajcency relation, NN becomes a digital topological group, which we call a digital topological subgroup of GG.

Note that a digital topological subgroup (N,κ|N)(N,\kappa|_{N}) of (G,κ)(G,\kappa) is not necessarily digital connected even if (G,κ)(G,\kappa) is κ\kappa-connected.

Let NN be a normal digital topological subgroup of a digital topological group GG. We define an adjacency relation on G/NG/N as follows.

Definition 4.4.

Let G=(G,κ)G=(G,\kappa) be a digital topological group and NN a normal digital topological subgroup, and let π:GG/N\pi:G\to G/N be the canonical surjection. We define an adjacency relation κ¯\bar{\kappa} on G/NG/N by declaring x¯κ¯y¯\bar{x}\leftrightarrow_{\bar{\kappa}}\bar{y} whenever there are xπ1(x¯)x\in\pi^{-1}(\bar{x}) and yπ1(y¯)y\in\pi^{-1}(\bar{y}) with xκyx\leftrightarrow_{\kappa}y in GG.

Proposition 4.5.

Let (G,κ)(G,\kappa) be an NPi\textrm{NP}_{i}-digital topological group, with an NPi\textrm{NP}_{i}-digital topological subgroup NN for some i{1,2}i\in\{1,2\}. Then the quotient group G/NG/N is an NPi\textrm{NP}_{i}-digital topological group with adjacency relation κ¯\bar{\kappa} as in Definition 4.4.

Proof.

We note that the projection map

π:(G,κ)(G/K,κ¯)\pi:(G,\kappa)\rightarrow(G/K,\bar{\kappa})

given by π(x)=x¯\pi(x)=\bar{x} is (κ,κ¯)(\kappa,\bar{\kappa})-continuous by Definition 4.4; that is, if xκyx\leftrightarrow_{\kappa}y in GG, then x¯κ¯y¯.\bar{x}\leftrightarrow_{\bar{\kappa}}\bar{y}.

Let μG:G×GG\mu_{G}:G\times G\rightarrow G and ιG:GG\iota_{G}:G\rightarrow G be the digital multiplication and the digital inverse, respectively, as in Definition 3.1. Define

μG/N:G/N×G/NG/N\mu_{G/N}:G/N\times G/N\rightarrow G/N

and

ιG/N:G/NG/N\iota_{G/N}:G/N\rightarrow G/N

by

μG/N(x¯,y¯)=xy¯\mu_{G/N}(\bar{x},\bar{y})=\overline{xy}

and

ιG/N(x¯)=x1¯,\iota_{G/N}(\bar{x})=\overline{x^{-1}},

respectively, for all x,yGx,y\in G. It can be seen that the following diagrams

G×G\textstyle{G\times G\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π×π\scriptstyle{\pi\times\pi}μG\scriptstyle{\mu_{G}}G\textstyle{G\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}and\textstyle{\mathrm{and}}G\textstyle{G\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}ιG\scriptstyle{\iota_{G}}G\textstyle{G\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}G/N×G/N\textstyle{G/N\times G/N\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μG/N\scriptstyle{\mu_{G/N}}G/N\textstyle{G/N}G/N\textstyle{G/N\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ιG/N\scriptstyle{\iota_{G/N}}G/N\textstyle{G/N}

are commutative. Moreover, if

(x¯1,y¯1)NPi(κ¯,κ¯)(x¯2,y¯2)(\bar{x}_{1},\bar{y}_{1})\leftrightarrow_{\textrm{NP}_{i}(\bar{\kappa},\bar{\kappa})}(\bar{x}_{2},\bar{y}_{2})

in G/N×G/NG/N\times G/N for some i{1,2}i\in\{1,2\}, and

z¯1κ¯z¯2\bar{z}_{1}\leftrightarrow_{\bar{\kappa}}\bar{z}_{2}

in G/NG/N, then we have

μG/N(x¯1,y¯1)=x1y1¯κ¯x2y2¯=μG/N(x¯2,y¯2)\mu_{G/N}(\bar{x}_{1},\bar{y}_{1})=\overline{x_{1}y_{1}}\leftrightarrow_{\bar{\kappa}}\overline{x_{2}y_{2}}=\mu_{G/N}(\bar{x}_{2},\bar{y}_{2})

and

ιG/N(z¯1)=z1¯κ¯z2¯=ιG/N(z¯2).\iota_{G/N}(\bar{z}_{1})=\overline{z_{1}}\leftrightarrow_{\bar{\kappa}}\overline{z_{2}}=\iota_{G/N}(\bar{z}_{2}).

Therefore, μG/N\mu_{G/N} and ιG/N\iota_{G/N} are (NPi(κ¯,κ¯),κ¯)(\textrm{NP}_{i}(\bar{\kappa},\bar{\kappa}),\bar{\kappa})-continuous and (κ¯,κ¯)(\bar{\kappa},\bar{\kappa})-continuous functions, respectively, for some i{1,2}i\in\{1,2\}, as required. ∎

We note that, when GG is a digital topological group embedded in (n,cu)(\mathbb{Z}^{n},c_{u}) and NN is a normal subgroup, the quotient group G/NG/N may not nicely take the form of a digital image in n\mathbb{Z}^{n} with cuc_{u} adjacency.

For example, let G(2,c1)G\subset(\mathbb{Z}^{2},c_{1}) be the simple closed curve of 16 points in Figure 3. We will label the points of GG as x0,x1,,x15x_{0},x_{1},\dots,x_{15} with xixi+1x_{i}\leftrightarrow x_{i+1} with subscripts read modulo 1616, so that GG is isomorphic as a group to 16\mathbb{Z}_{16}. Let N={x0,x4,x8,x12}N=\{x_{0},x_{4},x_{8},x_{12}\}, indicated by the circled points in Figure 3. Choosing {x¯0,x¯1,x¯2,x¯3}\{\bar{x}_{0},\bar{x}_{1},\bar{x}_{2},\bar{x}_{3}\} as coset representatives, the digital image (G/N,c¯1)(G/N,\bar{c}_{1}) is not naturally situated as a digital image in 2\mathbb{Z}^{2} with c1c_{1}-adjacency. (Though by Theorem 2.1, it will be isomorphic to some digital image in n\mathbb{Z}^{n} with cnc_{n}-adjacency. In fact it is a simple closed curve of 4 points.)

x0x_{0}(a)
x¯0\bar{x}_{0}(b)
Figure 3. (a) The digital topological group G(2,c1)G\subset(\mathbb{Z}^{2},c_{1}) isomorphic to 16\mathbb{Z}_{16}, with points from the subgroup N4N\cong\mathbb{Z}_{4} circled. (b) The quotient group G/NG/N, with adjacencies given by c¯1\bar{c}_{1}

For sets A,B(X,κ)A,B\subset(X,\kappa) in some digital image, we will say that AA and BB are adjacent if there exists points aAa\in A and bBb\in B with aba\leftrightarroweq b. In this case we write ABA\leftrightarroweq B.

Definition 4.6.

We call a (κ,λ)(\kappa,\lambda)-continuous function f:(X,κ)(Y,λ)f:(X,\kappa)\to(Y,\lambda) a digital open map when: if z,wf(G)z,w\in f(G) with zwz\leftrightarrow w, then f1(z)f1(w)f^{-1}(z)\leftrightarroweq f^{-1}(w).

Thus an open map is one for which f1f^{-1} preserves adjacencies in the same way that a continuous map does. The definition of digital open map is related to continuity of the preimage f1f^{-1} viewed as a multivalued map f1:Y2Y;f^{-1}:Y\to 2^{Y}; see [7, Theorems 2.4 and 2.7]. We use the term open map because it is required for Proposition 4.7, which in the classical theory requires that the map be open.

The following is the digital version of the first isomorphism theorem from group (or module) theory.

Theorem 4.7.

Let f:(G,κ)(H,λ)f:(G,\kappa)\rightarrow(H,\lambda) be a digital topological group homomorphism with kernel KK. If ff is a digital open map, then there exists a unique digital topological group isomorphism

F:G/KImfF:G/K\rightarrow{\rm Im}f

such that

Fπ(x)=f(x)F\circ\pi(x)=f(x)

for all xGx\in G, where

π:(G,κ)(G/N,κ¯)\pi:(G,\kappa)\rightarrow(G/N,\bar{\kappa})

is the (κ,κ¯)(\kappa,\bar{\kappa})-continuous function given by π(x)=x¯\pi(x)=\bar{x}.

Proof.

From the first isomorphism theorem from group theory, we see that there exists a unique isomorphism of groups

F:G/K\textstyle{F:G/K\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}ImfH\textstyle{{\rm Im}f\subseteq H}

such that the following diagram

G\textstyle{G\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}f\scriptstyle{f}ImfH\textstyle{{\rm Im}f\subseteq H}G/K\textstyle{G/K\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F\scriptstyle{F}

is commutative in the category of groups and group homomorphisms in the sense that

f(x)=Fπ(x)=F(x¯)f(x)=F\circ\pi(x)=F(\bar{x})

for all xGx\in G, where Imf\operatorname{Im}f is a digital topological subgroup of (H,λ)(H,\lambda). We need only show that FF is a digital isomorphism.

To show FF is continuous, take x¯κ¯y¯\bar{x}\leftrightarrow_{\bar{\kappa}}\bar{y} in (G/K,κ¯)(G/K,\bar{\kappa}). Then there are some xπ1(x)x\in\pi^{-1}(x) and yπ1(y)y\in\pi^{-1}(y) with xκyx\leftrightarrow_{\kappa}y. Since ff is continuous, we will have f(x)λf(y)f(x)\leftrightarrow_{\lambda}f(y). Thus we have

F(x¯)=f(x)λf(y)=F(y¯);F(\bar{x})=f(x)\leftrightarrow_{\lambda}f(y)=F(\bar{y});

and so FF is continuous.

Now to show F1F^{-1} is continuous, take f(x)λf(y)f(x)\leftrightarrow_{\lambda}f(y) in Imf\operatorname{Im}f, and we will show that

F1(f(x))κ¯F1(f(y))F^{-1}(f(x))\leftrightarrow_{\bar{\kappa}}F^{-1}(f(y))

in (G/K,κ¯)(G/K,\bar{\kappa}). Since F1f=πF^{-1}\circ f=\pi, we must show that

x¯κ¯y¯.\bar{x}\leftrightarrow_{\bar{\kappa}}\bar{y}.

Since ff is a digital open map and f(x)λf(y)f(x)\leftrightarrow_{\lambda}f(y), we have f1(f(x))f1(f(y))f^{-1}(f(x))\leftrightarroweq f^{-1}(f(y)), and thus there is some af1(f(x))a\in f^{-1}(f(x)) and bf1(f(y))b\in f^{-1}(f(y)) so that

x¯=π(x)=π(a)=a¯κ¯b¯=π(b)=π(y)=y¯\bar{x}=\pi(x)=\pi(a)=\bar{a}\leftrightarrow_{\bar{\kappa}}\bar{b}=\pi(b)=\pi(y)=\bar{y}

in (G/K,κ¯)(G/K,\bar{\kappa}) as required. ∎

As the following example shows, not all continuous homomorphisms of digital topological groups are open maps, and the conclusion of Theorem 4.7 may fail when the map is not open.

Example 4.8.

Let

X:={(1,0),(0,1),(1,0),(0,1)}X:=\{(1,0),(0,1),(-1,0),(0,-1)\}

be a digital image, pictured in Figure 4.

(1,0)(1,0)(0,1)(0,1)(1,0)(-1,0)(0,1)(0,-1)
Figure 4. The digital image XX in 2\mathbb{Z}^{2} from Example 4.8.

The digital image (X,c2)(X,c_{2}) is a simple cycle of 4 points, while (X,c1)(X,c_{1}) is a set of 4 never-adjacent points. Interpreting points of XX as complex numbers, the complex multiplication makes XX into a group isomorphic to 4\mathbb{Z}_{4}. It is easy to verify that this multiplication makes both (X,c1)(X,c_{1}) and (X,c2)(X,c_{2}) into NP1\textrm{NP}_{1}-digital topological groups.

Let f:(X,c1)(X,c2)f:(X,c_{1})\to(X,c_{2}) be the identity map. Then clearly f:XXf:X\to X is a continuous homomorphism, but it does not satisfy the open map condition. For example we have (1,0)c2(0,1)(1,0)\leftrightarrow_{c_{2}}(0,1), but f1(1,0)f^{-1}(1,0) and f1(0,1)f^{-1}(0,1) have no c1c_{1}-adjacent points.

In this case the conclusion of Proposition 4.7 will fail: the kernel of ff is trivial, so we have (X,c1)/kerf(X,c1)(X,c_{1})/\ker f\cong(X,c_{1}), while Im(f)(X,c2)\operatorname{Im}(f)\cong(X,c_{2}). But f:(X,c1)(X,c2)f:(X,c_{1})\to(X,c_{2}) is not a digital isomorphism.

The following theorem shows that a group homomorphism (not assumed to be continuous) of digital topological groups is automatically continuous provided that it is continuous near the identity element. This is analogous to the classical fact that a homomorphism of topological groups will be continuous if it is assumed to be continuous in a neighborhood of the identity element.

Theorem 4.9.

Let GG and HH be digital topological groups with identity elements eGe_{G} and eHe_{H}, respectively, and let f:GHf:G\to H be a group homomorphism with the property that: if xeGx\leftrightarrow e_{G}, then f(x)eHf(x)\leftrightarroweq e_{H}. Then ff is continuous and thus is a digital topological group homomorphism.

Proof.

Take x,yGx,y\in G, and we must show that f(x)f(y)f(x)\leftrightarroweq f(y) in HH. Since xyx\leftrightarrow y, we can apply the isomorphism μx1\mu_{x^{-1}} to obtain eGxy1e_{G}\leftrightarrow xy^{-1}. Then by our assumption on ff we have eHf(xy1)=f(x)f(y)1e_{H}\leftrightarroweq f(xy^{-1})=f(x)f(y)^{-1}. Then applying the isomorphism μf(y)\mu_{f(y)} gives f(y)f(x)f(y)\leftrightarrow f(x) as desired. ∎

We note that we can construct the category 𝒟𝒯𝒢\mathcal{DTG} whose object and morphism classes consist of digital topological groups and digital topological group homomorphisms, respectively.


5. Classification of NP2\textrm{NP}_{2}-digital topological groups

The NP2\textrm{NP}_{2}-continuity required for an NP2\textrm{NP}_{2}-digital topological group is very restrictive, and we can prove a complete classification of these groups. Recall that a cluster graph [16] is a graph which is a disjoint union of complete graphs.

In this section we show that a digital image is an NP2\textrm{NP}_{2}-digital topological group if and only if it is a regular cluster graph. The result is based on the following lemma:

Lemma 5.1.

If G:=(G,κ)G:=(G,\kappa) is a connected NP2\textrm{NP}_{2}-digital topological group, then every element of GG is adjacent or equal to the identity element eGe\in G.

Proof.

Let xGx\in G, and we will show that xx is adjacent or equal to ee. To obtain a contradiction, assume that xx is not adjacent or equal to ee, and let (x0,x1,,xn)(x_{0},x_{1},\dots,x_{n}) be a minimal path in GG from x0=ex_{0}=e to xn=xx_{n}=x with n2n\geq 2. Minimality of the path implies that x2x_{2} is not equal or adjacent to ee.

Since (e,x1)(e,x_{1}) is adjacent to (x1,x2)(x_{1},x_{2}) in (G×G,NP2(κ,κ))(G\times G,\textrm{NP}_{2}(\kappa,\kappa)), and the multiplication is NP2\textrm{NP}_{2}-continuous, we see that x1x_{1} is adjacent or equal to x1x2x_{1}x_{2}. Applying the isomorphism μx11\mu_{x_{1}^{-1}} shows that ee is adjacent or equal to x2x_{2}, which is a contradiction. ∎

The result above can be improved:

Theorem 5.2.

If GG is a connected NP2\textrm{NP}_{2}-digital topological group, then GG is a complete graph.

Proof.

Let x,yGx,y\in G, and we must show that xx is adjacent or equal to yy. Lemma 5.1 above shows that xx and yy must be adjacent to ee. Thus (e,x)(e,x) is NP2\textrm{NP}_{2}-adjacent to (y,e)(y,e), so multiplication shows that xx is adjacent to yy as desired. ∎

By Corollary 3.19, any digital topological group is a disjoint union of components, each isomorphic to GeG_{e}. Thus we obtain:

Corollary 5.3.

If GG is an NP2\textrm{NP}_{2}-digital topological group, then GG is a regular cluster graph.

We also have a converse to the above: any set which is a regular cluster graph can be given the structure of an NP2\textrm{NP}_{2}-digital topological group.

Theorem 5.4.

Let X:=(X,κ)X:=(X,\kappa) be a regular cluster graph. Then there are operations μX:X×XX\mu_{X}:X\times X\to X and ιX:XX\iota_{X}:X\to X which make XX into an NP2\textrm{NP}_{2}-digital topological group.

Proof.

Assume that XX consists of kk disjoint components, each a complete graph on nn vertices. We will construct coordinates for points of XX which label each point with a pair (i,j)n×k(i,j)\in\mathbb{Z}_{n}\times\mathbb{Z}_{k}, and show that, with respect to these coordinates, XX can be given the group structure of n×k\mathbb{Z}_{n}\times\mathbb{Z}_{k}.

Choose some arbitrary point of XX and label it as (0,0)(0,0). Each of the other points in the component of (0,0)(0,0) will be labeled as (i,0)(i,0) for ini\in\mathbb{Z}_{n}. Let X0,X1,,Xk1X_{0},X_{1},\dots,X_{k-1} be the components of XX, labeled so that (0,0)X0(0,0)\in X_{0}. For each jkj\in\mathbb{Z}_{k}, let gj:X0Xjg_{j}:X_{0}\to X_{j} be some graph isomorphism, with g0g_{0} being the identity map. Then for any jkj\in\mathbb{Z}_{k}, we label the points of XjX_{j} as

(i,j)=gj(i,0).(i,j)=g_{j}(i,0).

Note that since XX is a cluster graph, two points (i,j)(i,j) and (a,b)(a,b) in XX are adjacent in XX if and only if they are in the same component, which is to say

j=b.j=b.

Now we define the operations μX\mu_{X} and ιX\iota_{X} so that XX becomes an NP2\textrm{NP}_{2}-digital topological group with group structure isomorphic to n×k\mathbb{Z}_{n}\times\mathbb{Z}_{k}. We define:

μX((i,j),(a,b))=(i+a,j+b) and ιX(i,j)=(i,j).\mu_{X}((i,j),(a,b))=(i+a,j+b)\text{ and }\iota_{X}(i,j)=(-i,-j).

Clearly these operations make XX into a group. By Lemma 3.5 we need only show that μX\mu_{X} is NP2\textrm{NP}_{2}-continuous. Take ((i,j),(a,b))((i,j),(a,b)) to be NP2\textrm{NP}_{2}-adjacent to ((i,j),(a,b))((i^{\prime},j^{\prime}),(a^{\prime},b^{\prime})). This means that j=jj=j^{\prime} and b=bb=b^{\prime}. Then their products will be (i+a,j+b)(i+a,j+b) and (i+a,j+b)(i^{\prime}+a^{\prime},j^{\prime}+b^{\prime}) which are adjacent in XX because the second coordinates are equal. Thus μX\mu_{X} is NP2\textrm{NP}_{2}-continuous as desired. ∎



References

  • [1] J. A. Bondy and U. S. R. Murty, Graph theory, Graduate Texts in Mathematics, 244. Springer, New York, 2008.
  • [2] L. Boxer, Digitally continuous functions, Pattern Recognition Letters 15 (1994), 833-839.
  • [3] L. Boxer, A classical construction for the digital fundamental group, J. Math. Imaging Vis. 10 (1999), 51-62.
  • [4] L. Boxer, Properties of digital homotopy, J. Math. Imaging Vis. 22 (2005), 19-26.
  • [5] L. Boxer, Homotopy properties of sphere-like digital images, J. Math. Imaging Vis. 24 (2006), 167-175.
  • [6] L. Boxer, Generalized normal product adjacency in digital topology, Appl. Gen. Topol., 18(2) (2017), 401-427.
  • [7] L. Boxer and P. C. Staecker, Connectivity preserving multivalued functions in digital topology, Journal of Mathematical Imaging and Vision 55 (2016), 370–377.
  • [8] G. Chartrand and P. Zhang, Introduction to Graph Theory, McGraw-Hill Science, New York, 2005
  • [9] O. Ege and I. Karaca, Digital H-spaces, 3rd International Symposium on Computing in Science and Engineering, ISCSE 2013, 133-138.
  • [10] T. Husain, Introduction to topological groups, W. B. Saunders Co., Philadelphia, Pa.-London 1966.
  • [11] D.-W. Lee, Near-rings on digital Hopf groups, Appl. Algebra Engrg. Comm. Comput. 29(3) (2018), 261-282.
  • [12] D.-W. Lee, Digital H-spaces and actions in the pointed digital homotopy category, Appl. Algebra Engrg. Comm. Comput. 31(2) (2020), 149-169.
  • [13] G. Lupton, J. Oprea, N. Scoville, The Digital Hopf Construction, arxiv 2002.03027, 2020.
  • [14] D. Montgomery and L. Zippin, Topological Transformation Groups, Interscience Tracts in Pure and Applied Mathematics, Track 1, Interscience Publ. Inc. New York, 1955.
  • [15] A. Rosenfeld, Continuous functions on digital pictures, Pattern Recognition Letters 4 (1986), 177-184.
  • [16] R. Shamir, R. Sharan, and D. Tsur, Cluster graph modification problems, Discrete Appl. Math. 144 (2004), no. 1-2, 173–182.
  • [17] P. C. Staecker, Digital homotopy relations and digital homology theories, Appl. Gen. Topol. 22 (2021), no. 2, 223-250.
  • [18] D. B. West, Introduction to Graph Theory, Prentice Hall, Inc., Upper Saddle River, NJ, 1996.