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

Locally closed sets and submaximal spaces

R. Mohamadian Department of Mathematics, Shahid Chamran University of ahvaz, Ahvaz, Iran [email protected]
Abstract.

A topological space XX is called submaximal if every dense subset of XX is open. In this paper, we show that if βX\beta X, the Stone-Čech compactification of XX, is a submaximal space, then XX is a compact space and hence βX=X\beta X=X. We also prove that if υX\upsilon X, the Hewitt realcompactification of XX, is submaximal and first countable and XX is without isolated point, then XX is realcompact and hence υX=X\upsilon X=X. We observe that every submaximal Hausdorff space is pseudo-finite. It turns out that if υX\upsilon X is a submaximal space, then XX is a pseudo-finite μ\mu-compact space. An example is given which shows that XX may be submaximal but υX\upsilon X may not be submaximal. Given a topological space (X,𝒯)(X,{\mathcal{T}}), the collection of all locally closed subsets of XX forms a base for a topology on XX which is denotes by 𝒯l{\mathcal{T}_{l}}. We study some topological properties between (X,𝒯)(X,{\mathcal{T}}) and (X,𝒯l)(X,{\mathcal{T}_{l}}), such as we show that a) (X,𝒯l)(X,{\mathcal{T}_{l}}) is discrete if and only if (X,𝒯)(X,{\mathcal{T}}) is a TDT_{D}-space; b) (X,𝒯)(X,{\mathcal{T}}) is a locally indiscrete space if and only if 𝒯=𝒯l{\mathcal{T}}={\mathcal{T}_{l}}; c) (X,𝒯)(X,{\mathcal{T}}) is indiscrete space if and only if (X,𝒯l)(X,{\mathcal{T}_{l}}) is connected. We see that, in locally indiscrete spaces, the concepts of T0T_{0}, TDT_{D}, T12T_{\frac{1}{2}}, T1T_{1}, submaximal and discrete coincide. Finally, we prove that every clopen subspace of an lc-compact space is lc-compact.

Key words and phrases:
locally closed set, locally indiscrete space, TDT_{D}-space, submaximal space, Stone-Čech compactification, Hewitt realcompactification.
2010 Mathematics Subject Classification:
54G99, 54G05, 54G10, 54C30

1. Introduction

Throughout this paper we consider topological spaces on which no separation axioms are assumed unless explicity stated. The topology of a space is denoted by 𝒯{\mathcal{T}} and (X,𝒯)(X,{\mathcal{T}}) will be replaced by XX if there is no chance for confusion. For a subset AA of (X,𝒯)(X,{\mathcal{T}}), the closure, the interior, the boundary and the set of accumulation points of AA are denoted by cl𝒯(A){\rm cl}_{\mathcal{T}}(A) or clX(A){\rm cl}_{X}(A), int𝒯(A){\rm int}_{\mathcal{T}}(A) or intX(A){\rm int}_{X}(A), Fr𝒯(A){\rm Fr}_{\mathcal{T}}(A) or FrX(A){\rm Fr}_{X}(A) and AA^{\prime}, respectively. In places where there is no chance for confusion A¯\overline{A} and AA^{\circ} stands for cl𝒯(A){\rm cl}_{\mathcal{T}}(A) and int𝒯(A){\rm int}_{\mathcal{T}}(A), respectively. If AXA\subseteq X be open, closed or locally closed respect to the topology 𝒯{\mathcal{T}} on XX, then we write sometimes to avoid confusion, 𝒯{\mathcal{T}}-open, 𝒯{\mathcal{T}}-closed or 𝒯{\mathcal{T}}-locally closed, respectively. A subset AA of a topological space XX is called locally closed if for every xAx\in A there is an open set UXU\subseteq X such that xUx\in U and AUA\cap U is closed in UU. Since the intersection of two locally closed sets is locally closed, the family of 𝒯{\mathcal{T}}-locally closed sets forms a base for a finer topology 𝒯l{\mathcal{T}_{l}} on XX. For a Tychonoff (completely regular Hausdorff) space XX, βX\beta X is the Stone-Čech compactification and υX\upsilon X is the Hewitt realcompactification of XX. It is well known that XX is compact, pseudocompact or realcompact if and only if βX=X\beta X=X, βX=υX\beta X=\upsilon X or υX=X\upsilon X=X, respectively. For a Tychonoff space the symbol C(X)C(X) (resp. C(X)C^{*}(X)) denotes the ring of all continuous real valued (resp. all bounded continuous real valued) functions defined on XX. We say that a subspace SS of XX is CC^{*}-embedded in XX if every function in C(S)C^{*}(S) can be extended to a function in C(X)C^{*}(X). For fC(X)f\in C(X), the zero-set of ff is the set Z(f)={xX:f(x)=0}Z(f)=\{x\in X:f(x)=0\}. The set-theoretic complement of Z(f)Z(f) is denoted by coz(f)coz(f). The support of a function fC(X)f\in C(X) is the coz(f)¯\overline{coz(f)}. For any pβXp\in\beta X, the maximal ideal MpM^{p} (resp. the ideal OpO^{p}) is the set of all fC(X)f\in C(X) for which pclβXZ(f)p\in{\rm cl}_{\beta X}Z(f) (resp. pintβXclβXZ(f)p\in{\rm int}_{\beta X}{\rm cl}_{\beta X}Z(f)). More generally, for AβXA\subseteq\beta X, MAM^{A} (resp. OAO^{A}) is the intersection of all MpM^{p} (resp. OpO^{p}) with pAp\in A. It is clear that OAMAO^{A}\subseteq M^{A}.

A space XX is said to be a, a) door space if every set is either open or closed; b) T12T_{\frac{1}{2}}-space if every singleton is either open or closed; c) submaximal space if every dense subset is open; d) principal space, if every intersection of open sets is open, e) cc-principal space, if every countable intersection of open sets (i.e., GδG_{\delta}-set) is open, f) resolvable space if it has two disjoint dense subsets.

Tychonoff cc-principal spaces are the same PP-spaces, see Exercise 4J of [12]. A door space is submaximal and a submaximal space is T12T_{\frac{1}{2}}. The one-point compactification of any discrete space is submaximal. A nonempty resolvable space never submaximal. For example, \mathbb{R} and β\beta\mathbb{Q} are resolvable spaces. For more information about the locally closed sets, see [2, 8, 11], about the submaximal spaces, see [2, 8, 7, 1], about the door spaces, see [7, 9] and about the T12T_{\frac{1}{2}}-spaces, see [3]. For details about βX\beta X and υX\upsilon X, see [12, 18] and about other concepts of general topology, see [10, 19].

Section 2, of this paper is devoted to investigate some separation properties between (X,𝒯)(X,{\mathcal{T}}) and (X,𝒯l)(X,{\mathcal{T}_{l}}). Such as we show that (X,𝒯l)(X,{\mathcal{T}_{l}}) is discrete if and only if (X,𝒯)(X,{\mathcal{T}}) is a TDT_{D}-space. We show that, in this section, if βX\beta X is submaximal, then XX is compact and therefor, in this case, we conclude that X=βXX=\beta X. We also prove that if υX\upsilon X is submaximal and first countable and XX is without isolated point, then XX is realcompact. We observe that every submaximal Hausdorff space is pseudo-finite. It turns out that if υX\upsilon X is a submaximal space, then XX is a pseudo-finite μ\mu-compact space. In Section 3, we study and investigate the behavior locally indiscrete spaces. We observe that (X,𝒯)(X,{\mathcal{T}}) is a locally indiscrete space if and only if 𝒯=𝒯l{\mathcal{T}}={\mathcal{T}_{l}} if and only if every dense open subset of XX is regular open. In Section 4, we introduce some lc-properties such as lc-regular and lc-completely regular and compare them with the concepts regular and completely regular. We prove that every clopen subset of a lc-compact space is lc-compact.

2. Locally closed sets and 𝒯l{\mathcal{T}_{l}}-topology

A space XX is called a TDT_{D}-space if every singleton is locally closed. It is well known that the space XX is submaximal if and only if every subset of XX is locally closed, see Theorem 4.2 of [8] and also every subspace of a submaximal space is submaximal, see Theorem 1.1 of [8].

The proof of the following proposition is straightforward.

Proposition 2.1.

The following statements are equivalent, for a subset AA of the space XX.
a) AA is a locally closed set.
b) A=GBA=G\cap B, which GG is open and BB is closed.
c) A=HA¯A=H\cap\overline{A}, which HH is open.
d) A=EFA=E-F, which EE and FF are closed.
e) A¯A\overline{A}-A is a closed set.
f) A(A(XA¯))A\subseteq(A\cup(X-\overline{A}))^{\circ}.
g) A(XA¯)A\cup(X-\overline{A}) is an open set.

Remark 2.2.

a) Every closed (resp. open) set is locally closed.
b) Every dense locally closed set is open.
c) The intersection of finite number locally closed sets is locally closed.
d) Every intersection of locally closed sets may not be locally closed.
e) The complement of a locally closed set need not be locally closed. For example, A={1n:n=1,}A=\{\frac{1}{n}:n=1,\cdots\} is a locally closed set in \mathbb{R} while A\mathbb{R}-A is not locally closed.
f) The union of two locally closed sets need not be locally closed. For example, suppose that A=(,0]A=(-\infty,0] and B=n=1(1n+1,1n)B=\bigcup_{n=1}^{\infty}(\frac{1}{n+1},\frac{1}{n}) in \mathbb{R}. Then ABA\cup B is not locally closed.
g) The union of two completely separated locally closed sets is locally closed.
h) If AA is preopen, (that is, Aintcl(A)A\subseteq{\rm int}{\rm cl}(A)) then it is open if and only if it is locally closed.
i) If A¯\overline{A} is open, where AA is locally closed, then AA is open.
j) If AA is locally closed and BB is clopen, then AB¯\overline{A\cap B} is locally closed.
k) If AA and BB are locally closed, then the equal AB¯=A¯B¯\overline{A\cap B}=\overline{A}\cap\overline{B} may not be true. For example, let X={a,b}X=\{a,b\} and 𝒯={,{a},X}{\mathcal{T}}=\{\emptyset,\{a\},X\}. Now suppose that A={a}A=\{a\} and B={b}B=\{b\}.
l) Let XX be a topological space and YY a subspace of XX. The set AYA\subseteq Y is locally closed in YY if and only if A=YBA=Y\cap B, where BB is locally closed in XX.
m) In a locally compact Hausdorff space, any subset is locally closed if and only if it is a locally compact set, see Theorem 18.4 of [19] or Corollary 3.3.10 of [10].
n) Let X,YX,Y are topological spaces, the continuous mapping f:XYf:X\to Y is one-to-one and let the one element family {f}\{f\} separates points and closed sets. If XX is a T1T_{1}-space and f(X)f(X) is open subset of YY, then f(X)f(X) is a TDT_{D}-space, see Lemma 2.3.19 of [10].
o) A locally closed set AA of XX is dense in XX if and only if there is an open set GG of XX such that AGA\cap G is dense in XX.
p) Every discrete subset of a T1T_{1}-space is locally closed.
q) AXA\subseteq X is locally closed in XX if and only if it is locally closed in A¯\overline{A}.

Given a topological space (X,𝒯)(X,{\mathcal{T}}), the collection of all locally closed subsets of XX forms a base for a topology on XX which is denotes by 𝒯l{\mathcal{T}_{l}}. It is clear that 𝒯𝒯l{\mathcal{T}}\subseteq{\mathcal{T}_{l}} and in locally indiscrete spaces we have 𝒯=𝒯l{\mathcal{T}}={\mathcal{T}_{l}}, see Proposition 3.4. In the sequel, we study some topological properties between (X,𝒯)(X,{\mathcal{T}}) and (X,𝒯l)(X,{\mathcal{T}_{l}}).

Proposition 2.3.

(X,𝒯)(X,{\mathcal{T}}) is indiscrete if and only if (X,𝒯l)(X,{\mathcal{T}_{l}}) is a connected space.

Proof.

()(\Rightarrow) Let G𝒯G\in{\mathcal{T}} and let GX\emptyset\neq G\neq X. Clearly G𝒯lG\in{\mathcal{T}_{l}} and XG𝒯lX-G\in{\mathcal{T}_{l}} and therefore (X,𝒯l)(X,{\mathcal{T}_{l}}) is disconnected.
()(\Leftarrow) Suppose that X=GHX=G\cup H, which GG and HH are two nonempty disjoint 𝒯l{\mathcal{T}_{l}}-open sets and let xGx\in G. Hence there is a 𝒯{\mathcal{T}}-locally closed set AA such that xAGx\in A\subseteq G. Therefor A=Ucl𝒯AA=U\cap{\rm cl}_{\mathcal{T}}A, where U𝒯U\in{\mathcal{T}}. If UXU\neq X, then we are done. Assume that U=XU=X. This implies that A=cl𝒯AA={\rm cl}_{\mathcal{T}}A, that is, XAX-A is a 𝒯{\mathcal{T}}-open. Since HXA\emptyset\neq H\subseteq X-A, we infer that XAX-A\neq\emptyset. Now if XA=XX-A=X, then A=A=\emptyset which is not true. Consequently, XAX\emptyset\neq X-A\neq X is a 𝒯{\mathcal{T}}-open set. This complete the proof. ∎

Proposition 2.4.

(X,𝒯)(X,{\mathcal{T}}) is a T0T_{0}-space if and only if (X,𝒯l)(X,{\mathcal{T}_{l}}) is a T0T_{0}-space.

Proof.

()(\Rightarrow) It is trivial.
()(\Leftarrow) Let x,yXx,y\in X and xyx\neq y. Hence there is G𝒯lG\in{\mathcal{T}_{l}} such that xGx\in G and yGy\notin G. Let AA be 𝒯{\mathcal{T}}-locally closed set such that xAGx\in A\subseteq G. Therefor A=Ucl𝒯AA=U\cap{\rm cl}_{\mathcal{T}}A, where U𝒯U\in{\mathcal{T}}. If yUy\notin U, then we are done. If yUy\in U, then ycl𝒯Ay\notin{\rm cl}_{\mathcal{T}}A. Otherwise yAy\in A and so yGy\in G which is not true. Thus there is V𝒯V\in{\mathcal{T}} such that yVy\in V and VA=V\cap A=\emptyset. This consequence that xVx\notin V and this complete the proof. ∎

If (X,𝒯)(X,{\mathcal{T}}) is a T1T_{1}-space then (X,𝒯l)(X,{\mathcal{T}_{l}}) is discrete. The converse is not true. For example, let X={a,b}X=\{a,b\} and 𝒯={,{a},X}{\mathcal{T}}=\{\emptyset,\{a\},X\}. For the converse see the next proposition.

Proposition 2.5.

(X,𝒯)(X,{\mathcal{T}}) is a TDT_{D}-space if and only if (X,𝒯l)(X,{\mathcal{T}_{l}}) is discrete.

Proof.

()(\Rightarrow) It is trivial.
()(\Leftarrow) Let xXx\in X. Hence {x}𝒯l\{x\}\in{\mathcal{T}_{l}}, and therefore there is a 𝒯{\mathcal{T}}-locally closed set AA such that xA{x}x\in A\subseteq\{x\}. It implies that A={x}A=\{x\}, i.e., (X,𝒯)(X,{\mathcal{T}}) is a TDT_{D}-space. ∎

Remark 2.6.

Every TDT_{D}-space is T0T_{0}. The converse is false. For example, we consider the topology 𝒯={(a,):a}{,}{\mathcal{T}}=\{(a,\infty):a\in\mathbb{R}\}\cup\{\emptyset,\mathbb{R}\} on \mathbb{R}.

If 𝒯{\mathcal{T}} be the principal topology on XX, then every element xXx\in X has a minimal open neighborhood Mx={G𝒯:xG}M_{x}=\bigcap\{G\in{\mathcal{T}}:x\in G\}. For details about principal topology, see [17].

Lemma 2.7.

Every principal T0T_{0}-space is TDT_{D}.

Proof.

Suppose that xXx\in X. If {x}¯={x}\overline{\{x\}}=\{x\}, then we are done. Hence, assume that xx0{x}¯x\neq x_{0}\in\overline{\{x\}}. Since there is no an open set GG such that x0Gx_{0}\in G and xGx\notin G we infer that there is an open set HH such that xHx\in H and x0Hx_{0}\notin H. Now let U=MxU=M_{x} be the minimal open neighborhood of xx. We claim that {x}=U{x}¯\{x\}=U\cap\overline{\{x\}}. Assume that tU{x}¯t\in U\cap\overline{\{x\}}. We show that t=xt=x. If not, then two cases occur. In the first case, there is an open set VV such that tVt\in V and xVx\notin V, which is not true, for t{x}¯t\in\overline{\{x\}}. In the second case, there is an open set WW such that tWt\notin W and xWx\in W. Now since tUt\in U, we conclude that tWt\in W which is not true. Consequently, t=xt=x and this implies that {x}\{x\} is locally closed set. ∎

Remark 2.8.

a) If (X,𝒯)(X,{\mathcal{T}}) is a principal space, then every intersection of locally closed sets is locally closed.
b) If (X,𝒯)(X,{\mathcal{T}}) is a principal space, then (X,𝒯l)(X,{\mathcal{T}_{l}}) is principal. The converse is false. For example, (,𝒯l)(\mathbb{R},{\mathcal{T}_{l}}) is principal but (,𝒯u)(\mathbb{R},{\mathcal{T}_{u}}), where 𝒯u{\mathcal{T}_{u}} denotes usual topology on \mathbb{R}, is not principal.

Example 2.9.

A continuous image of a TDT_{D}-space need not be TDT_{D}-space. Let X=(0,1)X=(0,1) with usual topology and Y=(1,)Y=(1,\infty) with 𝒯Y={YG:G𝒯}{\mathcal{T}_{Y}}=\{Y\cap G:G\in{\mathcal{T}}\}, where 𝒯={(a,):a}{,}{\mathcal{T}}=\{(a,\infty):a\in\mathbb{R}\}\cup\{\emptyset,\mathbb{R}\} on \mathbb{R}. We define f:XYf:X\to Y by f(x)=1xf(x)=\frac{1}{x}. Then ff is a one-to-one and onto which is continuous. To prove continuity, suppose that HYH\subseteq Y be open and f(x0)Hf(x_{0})\in H, where x0Xx_{0}\in X. There is 0<r0<r\in\mathbb{R} such that r<1x0r<\frac{1}{x_{0}} and H=(r,)H=(r,\infty). We put G=(s,1r)G=(s,\frac{1}{r}), where s<x0s<x_{0}. It is clear that f(G)Hf(G)\subseteq H. Note that XX is a TDT_{D}-space while YY is not a TDT_{D}-space.

If αΛXα\prod_{\alpha\in\Lambda}X_{\alpha} is a TDT_{D}-space, then XαX_{\alpha} is a TDT_{D}-space for each αΛ\alpha\in\Lambda. The converse is true if II is finite. In the next example we see that an arbitrary product of TDT_{D}-spaces need not be TDT_{D}.

Example 2.10.

Let Xn={a,b}X_{n}=\{a,b\} and 𝒯n={,{a},X}{\mathcal{T}_{n}}=\{\emptyset,\{a\},X\}, for any nn\in\mathbb{N} and let X=nXnX=\prod_{n\in\mathbb{N}}X_{n}. Suppose that x=(xn)Xx=(x_{n})\in X, where xn=ax_{n}=a for every nn\in\mathbb{N}. It is clear that xx belongs to every nonempty open set in XX. We claim that {x}\{x\} is not a locally closed set. Otherwise, there is an open set GG and a closed set FF such that {x}=GF\{x\}=G\cap F. If FXF\neq X, then xF(XF)x\in F\cap(X-F) which is not true. If F=XF=X, then {x}=G\{x\}=G. But G=nGnG=\prod_{n\in\mathbb{N}}G_{n}, where Gn=XnG_{n}=X_{n}, for every nIn\notin I, which II\subseteq\mathbb{N} is finite and this is impossible. It is consequence that {x}\{x\} is not locally closed.

Lemma 2.11.

X=αΛXαX=\prod_{\alpha\in\Lambda}X_{\alpha} is a cc-principal space if and only if all but finitely many of the XαX_{\alpha} are trivial spaces.

Proof.

It is straightforward. ∎

In the sequel, by |A||A| we mean the cardinality of AA for every set AA.

Remark 2.12.

Let X=αΛXαX=\prod_{\alpha\in\Lambda}X_{\alpha} be a cc-principal space and |I|0|I|\leq\aleph_{0}. Then XX is a TDT_{D}-space if and only if each factor space is a TDT_{D}-space, see part (b) of Exercise 2.3.B of [10].

Every T12T_{\frac{1}{2}}-space is TDT_{D}. The converse is not true. For example, let X=X=\mathbb{N} and 𝒯={En:n=1,}{}{\mathcal{T}}=\{E_{n}:n=1,\cdots\}\cup\{\emptyset\}, where En={n,n+1,}E_{n}=\{n,n+1,\cdots\}, for any nn\in\mathbb{N}. Since {n}=En(En+1)\{n\}=E_{n}\cap(\mathbb{N}-E_{n+1}), we infer that \mathbb{N} is a TDT_{D}-space. It is clear that the set {2}\{2\} neither open nor closed, so \mathbb{N} is not a T12T_{\frac{1}{2}}-space.

Recall that union of two locally closed sets may not be locally closed. In the next example, we see that even if we add a cluster point to a locally closed set, the resulting set may not be locally closed. In other words, if AA is locally closed and xA¯Ax\in\overline{A}-A, then A{x}A\cup\{x\} need not be locally closed.

Example 2.13.

Suppose that X=X=\mathbb{R}, and for any kk\in\mathbb{N}, let Ak={n+1kn+1:n=1,}A_{k}=\{\frac{n+1}{kn+1}:n=1,\cdots\}. We consider A=k=1AkA=\bigcup_{k=1}^{\infty}A_{k} and we put B=A{1,12,}B=A-\{1,\frac{1}{2},\cdots\}. Then B¯B={0,1,12,}\overline{B}-B=\{0,1,\frac{1}{2},\cdots\} is closed and hence BB is locally closed in \mathbb{R}. Now assume that C=B{0}C=B\cup\{0\}. Clearly, B¯=C¯\overline{B}=\overline{C}. Note that C¯C={1,12,}\overline{C}-C=\{1,\frac{1}{2},\cdots\} is not closed set in \mathbb{R} and hence CC is not locally closed.

Example 2.14.

An arbitrary product of Tychonoff, compact and submaximal spaces need not be submaximal. For example, assume that Xn=Xn{σn}X_{n}^{*}=X_{n}\cup\{\sigma_{n}\} be a one-point compactification of a discrete space XnX_{n}, for every nn\in\mathbb{N}. Suppose that X=nXnX=\prod_{n\in\mathbb{N}}X_{n}^{*} and suppose that A=nXnA=\prod_{n\in\mathbb{N}}X_{n}, then AA is dense in XX which is not open.

Proposition 2.15.

Let XX be a submaximal space and A,BXA,B\subseteq X with AB=A\cap B=\emptyset. Then A¯B¯\overline{A}\cap\overline{B} is discrete.

Proof.

Suppose that xA¯B¯x\in\overline{A}\cap\overline{B}. Hence A{x}A\cup\{x\} is locally closed in A¯\overline{A}. Therefor there is an open set GG in A¯\overline{A} such that A{x}=GclA¯(A{x})A\cup\{x\}=G\cap{\rm cl}_{\overline{A}}(A\cup\{x\}). Let UU be open in XX such that G=UA¯G=U\cap\overline{A}. Note that clA¯(A{x})=clX(A{x})A¯=A¯A¯=A¯{\rm cl}_{\overline{A}}(A\cup\{x\})={\rm cl}_{X}(A\cup\{x\})\cap\overline{A}=\overline{A}\cap\overline{A}=\overline{A}. This implies that A{x}=UA¯A\cup\{x\}=U\cap\overline{A}. Similarly, there is an open set VV in XX such that B{x}=VB¯B\cup\{x\}=V\cap\overline{B}. Now {x}=(A{x})(B{x})=(UA¯)(VB¯)=(UV)(A¯B¯)\{x\}=(A\cup\{x\})\cap(B\cup\{x\})=(U\cap\overline{A})\cap(V\cap\overline{B})=(U\cap V)\cap(\overline{A}\cap\overline{B}). This consequence that xx is an isolated point of A¯B¯\overline{A}\cap\overline{B}. ∎

If XX is a submaximal space and AXA\subseteq X, then Fr(A){\rm Fr}(A) is a discrete subset of XX. Also if XX is a submaximal space and AA is a discrete subset of XX, then A¯\overline{A^{\prime}} is discrete and if in addition XX is a T1T_{1}-space, then AA^{\prime} is discrete.

Theorem 2.16.

If βX\beta X is a submaximal space, then XX is compact. In this case βX=X\beta X=X.

Proof.

Since βX\beta X is countably compact, by Theorem 4.20 of [1], βX\beta X is a finite disjoint union of one-point compactification of some discrete spaces. Suppose that βX=i=1nXi\beta X=\bigcup_{i=1}^{n}X^{*}_{i}, where XiX^{*}_{i} is the one-point compactification of discrete space XiX_{i}, for every i=1,,ni=1,\cdots,n. Assume that Xi=Xi{σi}X^{*}_{i}=X_{i}\cup\{\sigma_{i}\}, Y=i=1nXiY=\bigcup_{i=1}^{n}X_{i} and A={σ1,,σn}A=\{\sigma_{1},\cdots,\sigma_{n}\}. Clearly, YY is discrete and hence YXY\subseteq X. It is sufficient to show that AXA\subseteq X. On the contrary suppose that A(βXX)=IA\cap(\beta X-X)=I is a nonempty set. Then D=iIXiD=\bigcup_{i\in I}X_{i} is discrete and X=D(iIXi)X=D\cup(\bigcup_{i\notin I}X^{*}_{i}). One can easily see that DD is CC^{*}-embedded in XX and consequently is CC^{*}-embedded in βX\beta X. Hence, clβXD=βD{\rm cl}_{\beta X}D=\beta D and therefor βDiIXi\beta D\subseteq\bigcup_{i\in I}X^{*}_{i}. This implies that |βD||iIXi|=iI|Xi|=|D||\beta D|\leq|\bigcup_{i\in I}X^{*}_{i}|=\sum_{i\in I}|X_{i}|=|D| which is contradiction, for by Theorem 9.2 of [12] we have |βD|=22|D||\beta D|=2^{2^{|D|}}. ∎

The converse of the above theorem is not true, in general. For example, X=βX=\beta\mathbb{Q} is compact, while βX=β\beta X=\beta\mathbb{Q} is not a submaximal space. Also, if XX is a submaximal space, then βX\beta X may not be submaximal. For example, \mathbb{N} is submaximal but β\beta\mathbb{N} is not submaximal.

Corollary 2.17.

Let υX\upsilon X be a submaximal space. The following statements are equivalent.
a) XX is compact.
b) XX is σ\sigma-compact.
c) XX is countably compact.
d) XX is pseudocompact.

Proof.

(ab)(a\Rightarrow b) and (cd)(c\Rightarrow d) are trivial.
(bc)(b\Rightarrow c) Since XX is union a finite number of compact spaces, then it is countably compact.
(ca)(c\Rightarrow a) If XX is a pseudocompact space, then υX=βX\upsilon X=\beta X and by hypothesis βX\beta X is submaximal. Now by Theorem 2.16 we conclude that XX is compact. ∎

Lemma 2.18.

If XX is a dense countably compact set in TT and TT is a Hausdorff submaximal space, then X=YX=Y.

Proof.

Since XX is a submaximal space, then by Theorem 4.21 of [1], XX is compact and since TT is Hausdorff, we infer that XX is closed in TT. Therefor, by density of XX we have X=TX=T. ∎

By the above lemma the proof of the next corollary is obvious.

Corollary 2.19.

If υX\upsilon X is submaximal and XX is countably compact, then XX is realcompact.

Remark 2.20.

We denote the set of all isolated points of space XX with I(X)I(X). If XX is open in TT, then I(X)I(T)I(X)\subseteq I(T) and if XX is dense in TT, then I(T)I(X)I(T)\subseteq I(X). Hence, if XX is an open dense set in TT, then I(X)=I(T)I(X)=I(T).

Theorem 2.21.

Suppose that TT is a first countable, Hausdorff and submaximal space and XX is a dense set in TT and I(X)=I(X)=\emptyset. Then X=TX=T.

Proof.

Let pTXp\in T-X. Hence, there is an infinite sequence (xn)(x_{n}) contained in XX which converges to pp. Put A={xn:n=1,}{p}A=\{x_{n}:n=1,\cdots\}\cup\{p\}. Let NN be a copy of \mathbb{N} contained in AA. Since AA is compact, then NN has an accumulation point in AA, namely x0x_{0}. We claim that TN¯=T\overline{T-N}=T. To see this, let aTa\in T and aGTa\in G\subseteq T, which GG is an open set in TT. We must show that G(TN)G\cap(T-N)\neq\emptyset. If not, then aGNa\in G\subseteq N and hence there is an open set HH in TT such that HN={a}H\cap N=\{a\}. It is clear that HG={a}H\cap G=\{a\}, that is aI(T)a\in I(T) which is contradiction. Now one can easily see that x0TNx_{0}\in T-N while x0(TN)x_{0}\notin(T-N)^{\circ}, that is TNT-N is not open, so it contradicts the submaximality of TT. ∎

Corollary 2.22.

Let XX be a compact submaximal space. Then |I(X)||I(X)| is finite if and only if XX is finite.

Proof.

Suppose that |I(X)||I(X)| is finite. If XI(X)X-I(X) is infinite, then similar to the proof of the Theorem 2.21, we can show that XX is not submaximal, which is a contradiction. Hence XI(X)X-I(X) is finite and so XX is finite. ∎

Corollary 2.23.

If υX\upsilon X is submaximal and first countable and I(X)=I(X)=\emptyset, then X=υXX=\upsilon X i.e., XX is realcompact.

Proof.

By Theorem 2.21 is clear. ∎

A space XX is called a pseudo-finite (resp. pseudo-discrete) space if every compact subspace of XX is finite (resp. has finite interior). We say that XX is real pseudo-finite, if υX\upsilon X is pseudo-finite. Every pseudo-finite space is pseudo-discrete, but not conversely. For example, we consider the space \mathbb{Q} of rational numbers.

Example 2.24.

The space Σ\Sigma is pseudo-finite. To see this, let FΣF\subseteq\Sigma be a compact subspace. If σF\sigma\notin F, then FF is discrete and hence it is finite. If σF\sigma\in F and FF is infinite, then assume that A={x1,x2,}A=\{x_{1},x_{2},\cdots\} be an infinite subset of FF and put G=({x1}){σ}G=(\mathbb{N}-\{x_{1}\})\cup\{\sigma\}. Now 𝒞={{xn}:n=1,}{G}{\mathcal{C}}=\{\{x_{n}\}:n=1,\cdots\}\cup\{G\} be an open cover for FF which has not a finite subcover. This is a contradiction.

Proposition 2.25.

Every submaximal Hausdorff space is pseudo-finite.

Proof.

Let XX be a submaximal Hausdorff space and let YY be a compact subspace of XX. If YY be infinite, then contains a copy of \mathbb{N}, namely NN. Suppose that x0Nx_{0}\in N^{\prime}, where x0Yx_{0}\in Y. Now similar to the proof of the Theorem 2.21, it is observe that XNX-N is dense in XX but it is not an open set in XX, so it contradicts the submaximality of XX. ∎

A pseudo-finite Tychonoff space may be not submaximal. For example, \mathbb{Q} is pseudo-finite but not submaximal. See also the following example.

Example 2.26.

We consider the space XX in Example 2 of [16]. This example shows that the Tychonoff space XX is an infinite countably compact subset of β\beta\mathbb{N} which is also pseudo-finite. We claim that XX is not submaximal. Otherwise, by Theorem 4.20 of [1], XX must be a compact space. In this case, since XX is pseudo-finite, it must be finite, which is not true.

CK(X)C_{K}(X) (resp. Cψ(X)C_{\psi}(X)) is the family of all fC(X)f\in C(X) with compact (resp. pseudocompact) support. Also C(X)C_{\infty}(X) is a subcollection of C(X)C^{*}(X) consisting of all functions vanishing at infinity, that is all fC(X)f\in C(X) for which {xX:|f(x)|1n}\{x\in X:|f(x)|\geq\frac{1}{n}\} is compact for every nn\in\mathbb{N}. For the convenience of readers, some special ideals are listed below.

a) CF(X)=OβXI(X)C_{F}(X)=O^{\beta X\setminus I(X)} is the intersection of all essential ideals in C(X)C(X). Recall that CF(X)C_{F}(X) is the socle of C(X)C(X) and it is well known that CF(X)={fC(X):coz(f)is finite}C_{F}(X)=\{f\in C(X):coz(f)~{}\mbox{is finite}\}, see Proposition 3.3 of [14].

b) CK(X)=OβXXC_{K}(X)=O^{\beta X\setminus X} is the intersection of all free ideals in C(X)C(X), see 7E of [12].

c) C(X)=MβXXC_{\infty}(X)=M^{*\beta X\setminus X} is the intersection of all free maximal ideals in C(X)C^{*}(X).

d) Cψ(X)=MβXυXC_{\psi}(X)=M^{\beta X\setminus\upsilon X} is the intersection of all hyper-real maximal ideals in C(X)C(X).

e) Iψ(X)=MβXXI_{\psi}(X)=M^{\beta X\setminus X} is the intersection of all free maximal ideals in C(X)C(X).

One can easily see that CF(X)CK(X)Iψ(X)C_{F}(X)\subseteq C_{K}(X)\subseteq I_{\psi}(X). In Theorem 8.19 of [12], it is shown that if XX is realcompact, then CK(X)=Iψ(X)C_{K}(X)=I_{\psi}(X). In part (a) of Theorem 4.5 of [4], it is proved that XX is a pseudo-discrete if and only if CF(X)=CK(X)C_{F}(X)=C_{K}(X).

In [13] a space XX is called:
a) a μ\mu-compact if CK(X)=Iψ(X)C_{K}(X)=I_{\psi}(X).
b) an η\eta-compact if Cψ(X)=Iψ(X)C_{\psi}(X)=I_{\psi}(X).
c) a ψ\psi-compact if CK(X)=Cψ(X)C_{K}(X)=C_{\psi}(X).
d) an \infty-compact if CK(X)=C(X)C_{K}(X)=C_{\infty}(X).

In [15] a space XX is called an ii-compact if C(X)=Iψ(X)C_{\infty}(X)=I_{\psi}(X).

Proposition 2.27.

If υX\upsilon X is a submaximal space, then XX is a pseudo-finite μ\mu-compact space.

Proof.

By Corollary 2.25, XX is pseudo-finite. Since XX is real pseudo-finite, by Theorem 3.8 of [14], we infer that CK(X)=Iψ(X)C_{K}(X)=I_{\psi}(X), that is, XX is a μ\mu-compact space. ∎

The converse of the above proposition is not true, in general. For example, \mathbb{Q} is a pseudo-finite μ\mu-compact space but υ=\upsilon\mathbb{Q}=\mathbb{Q} is not submaximal.
By part (b) of Theorem 4.5 of [4], the proof of the result is obvious.

Corollary 2.28.

Let XX be a submaximal Hausdorff space and |I(X)|<0|I(X)|<\aleph_{0}. Then CK(X)=C(X)C_{K}(X)=C_{\infty}(X).

Proposition 2.29.

Let υX\upsilon X be a submaximal space and |I(X)|<0|I(X)|<\aleph_{0}. Then the following statement are hold.
a) XX is a ii-compact.
b) XX is a \infty-compact.
c) XX is compact if and only if it is a realcompact space.

Proof.

The proof of parts (a) and (b) and implication (\Rightarrow) of (c) is clear. For the other side of (c), since XX is ii-compact, then C(X)=Iψ(X)C_{\infty}(X)=I_{\psi}(X). Since XX is realcompact, we infer that Cψ(X)=Iψ(X)=CK(X)C_{\psi}(X)=I_{\psi}(X)=C_{K}(X), that is, XX is η\eta-compact and ψ\psi-compact. Hence C(X)=Cψ(X)C_{\infty}(X)=C_{\psi}(X) and by Corollary 2.5 of [5] we conclude that XX is compact. ∎

Similar to Theorem 2.16, the question arises that if υX\upsilon X is a submaximal space, is XX realcompact?. So far, we have not been able to answer this question, in general. Therefor we express it as a question.

Question: If υX\upsilon X is a submaximal space, is XX realcompact?

The space \mathbb{R} is a realcompact space which is not submaximal. See the following example for an example of a space that is submaximal but not realcompact. This example, also shows that XX may be submaximal but υX\upsilon X may not be submaximal.

Example 2.30.

We consider the space Ψ{\mathbb{\Psi}} in 5I of [12]. By 5I. 5, the space Ψ{\mathbb{\Psi}} is pseudocompact which it is not a realcompact space. We now show that it is submaximal. Suppose that B¯=Ψ\overline{B}={\mathbb{\Psi}}. Since B\mathbb{N}\subseteq B, we infer that B=AB=\mathbb{N}\cup A, where A={ωE:E}A=\{\omega_{E}:E\in{\mathcal{F}\subseteq{\mathcal{E}}}\}. Assume that ωEB\omega_{E}\in B, which EE\in{\mathcal{F}}. Put G=D{ωE}G=D\cup\{\omega_{E}\}, where DD containing all but a finite number of points of EE. Then GG is an open set contains ωE\omega_{E} and GBG\subseteq B. This implies that ωEB\omega_{E}\in B^{\circ}, that is BB open and consequence that Ψ{\mathbb{\Psi}} is submaximal. Furthermore, since Ψ{\mathbb{\Psi}} is not compact, by Theorem 2.16, βΨ\beta{\mathbb{\Psi}} is not submaximal. Now υΨ=βΨ\upsilon{\mathbb{\Psi}}=\beta{\mathbb{\Psi}} shows that υΨ\upsilon{\mathbb{\Psi}} is not a submaximal space.

It is possible that XX is a locally compact and pseudocompact space but υX\upsilon X is not submaximal. See the next example.

Example 2.31.

We consider the space 𝕎=W(ω1)={σ:σ<ω1}\mathbb{W}=W(\omega_{1})=\{\sigma:\sigma<\omega_{1}\} of all countable ordinals and 𝕎=W(ω1+1)={σ:σω1}\mathbb{W}^{*}=W(\omega_{1}+1)=\{\sigma:\sigma\leq\omega_{1}\}, where ω1\omega_{1} denoting the first uncountable ordinal. It is well known that 𝕎\mathbb{W} is a pseudocompact locally compact space which neither compact nor realcompact. Clearly, υ𝕎=β𝕎=𝕎\upsilon\mathbb{W}=\beta\mathbb{W}=\mathbb{W}^{*}. Since 𝕎\mathbb{W} is not compact, then by Theorem 2.16, υ𝕎\upsilon\mathbb{W} is not submaximal. For a direct proof let A={ω0,2ω0,3ω0.}A=\{\omega_{0},2\omega_{0},3\omega_{0}.\cdots\}, where ω0\omega_{0} is the first infinite ordinal and let B=𝕎AB=\mathbb{W}^{*}-A. One can easily check that BB is dense in υ𝕎\upsilon\mathbb{W}. But BB is not open in υ𝕎\upsilon\mathbb{W}, for ω02B\omega_{0}^{2}\in B while ω02B\omega_{0}^{2}\notin B^{\circ}.

3. locally indiscrete spaces

A space XX is called locally indiscrete if every open set is closed or equivalently if every closed set is open. Every discrete space is locally indiscrete. For another nontrivial example, let RR be a principal ideal ring. Then the space Min(R){\rm Min}(R) with Zariski topology is a locally indiscrete space.

Proposition 3.1.

For a topological space XX the following conditions are equivalent.
a) XX is locally indiscrete.
b) Every subset of XX is preopen.
c) Every singleton in XX is preopen.
d) Every closed subset of XX is preopen.
e) Every locally closed subset of XX is open.
f) Every locally closed subset of XX is closed.
g) The closure of every locally closed set is open.
h) Every dense open subset of XX is regular open.

Proof.

All implications are straightforward. We only show (ha)(h\Rightarrow a). Suppose that AA is an open set in XX and consider B=A(XA)B=A\cup(X-A)^{\circ}. Then BB is open in XX. Furthermore, B¯=A¯(XA)¯=A¯(XA¯)¯=A¯(XA¯)=X\overline{B}=\overline{A}\cup\overline{(X-A)^{\circ}}=\overline{A}\cup\overline{(X-\overline{A})}=\overline{A}\cup(X-\overline{A}^{\circ})=X. Hence, by hypothesis BB is regular open and therefore we have X=B¯BX=\overline{B}^{\circ}\subseteq B. It implies that X=A(XA¯)X=A\cup(X-\overline{A}) and thus A¯(XA)=\overline{A}\cap(X-A)=\emptyset. This consequence that A¯A\overline{A}\subseteq A, that is, AA is closed and we are done. ∎

Proposition 3.2.

For a topological T1T_{1}-space XX the following conditions are equivalent.
a) XX is discrete.
b) XX is locally indiscrete.
c) Every open set is regular open.

Proof.

It is straightforward. ∎

To see the definition of the concepts given in the next remark, see [12, 10]. Foe details about PFP_{F}-spaces, see [6].

Remark 3.3.

a) A locally indiscrete space need not be discrete.
b) If XX is a T12T_{\frac{1}{2}}-space, then it is locally indiscrete if and only if it is discrete.
c) Every locally indiscrete space is a strongly zero-dimensional space. The converse is not true. For example, A={0,1,12,}A=\{0,1,\frac{1}{2},\cdots\} is a strongly zero-dimensional subspace of \mathbb{R} while it is not locally indiscrete.
d) Every locally indiscrete space is a PFP_{F}-space. The converse is not true. For example, the space Σ\Sigma of 4M in [12] is a PFP_{F}-space which is not locally indiscrete.
e) A locally indiscrete space need not be submaximal. For instance, let X={a,b,c}X=\{a,b,c\} and 𝒯={,{a},{b,c},X}{\mathcal{T}}=\{\emptyset,\{a\},\{b,c\},X\}.
f) A submaximal space need not be a locally indiscrete space. For example we consider the space Σ\Sigma of 4M in [12].
g) Every locally indiscrete space is a PP-space. The converse is false. For example, the space SS in 4.N of [12] is a PP-space which is not locally indiscrete. h) Every locally indiscrete space is a extremally disconnected and hence is basically disconnected. The converse is false. For example, the space SS in 4.N of [12] is a basically disconnected which is not locally indiscrete and the space Σ\Sigma is extremally disconnected which is not locally indiscrete.

Proposition 3.4.

The space (X,𝒯)(X,{\mathcal{T}}) is a locally indiscrete space if and only if 𝒯=𝒯l{\mathcal{T}}={\mathcal{T}_{l}}.

Proof.

()(\Rightarrow) Let G𝒯lG\in{\mathcal{T}_{l}}. Then G=αΛAαG=\bigcup_{\alpha\in\Lambda}A_{\alpha}, where AαA_{\alpha} is 𝒯{\mathcal{T}}-locally closed, for any αΛ\alpha\in\Lambda. By hypothesis, AαA_{\alpha} is 𝒯{\mathcal{T}}-open, hence G𝒯G\in{\mathcal{T}} and we are done.
()(\Leftarrow) Let AA is a 𝒯{\mathcal{T}}-locally closed subset of XX. Hence A𝒯lA\in{\mathcal{T}_{l}} and therefore A𝒯A\in{\mathcal{T}}. This shows that XX is a locally indiscrete space. ∎

Proposition 3.5.

Let XX be a locally indiscrete space. The following conditions are equivalent.
a) XX is a T1T_{1}-space.
b) XX is a T12T_{\frac{1}{2}}-space.
c) XX is a TDT_{D}-space.
d) XX is a T0T_{0}-space.
e) XX is a submaximal space.
f) XX is a discrete space.

Proof.

All implications are obvious. We only show (da)(d\Rightarrow a). Let xXx\in X and on the contrary suppose that y{x}¯y\in\overline{\{x\}} and yxy\neq x. If there is an open set GG such that yGy\in G and xGx\notin G, then G{x}G\cap\{x\}\neq\emptyset which is not true. If there is an open set HH such that xHx\in H and yHy\notin H, then yXHy\in X-H and since XHX-H is open we infer that (XH){x}(X-H)\cap\{x\}\neq\emptyset and this is not true. Therefor XX is a T1T_{1}-space and we are through. ∎

4. lc-properties

In this section, by using the locally closed sets, we introduce some separation axioms. For more details about lc-continuous functions, see [11]. We begin with the following definition.

Definition 4.1.

A space XX is called:
a) lc-regular if for each locally closed set AA and for each point xAx\notin A, there are disjoint open sets UU and VV with xUx\in U and AVA\subseteq V.
b) lc-completely regular if for each locally closed set AA and for each point xAx\notin A, there exists a continuous function f:X[0,1]f:X\to[0,1] such that f(x)=0f(x)=0 and f(A)=1f(A)=1.
c) lc-normal if for every two disjoint locally closed sets AA and BB, there are disjoint open sets UU and VV with AUA\subseteq U and BVB\subseteq V.

Every lc-regular (resp. lc-completely regular, lc-normal) space is a regular (resp. completely regular, normal) space. The converse is hold in locally indiscrete spaces, but is not true, in general. See the following example.

Example 4.2.

We consider X=X=\mathbb{R} with usual topology.
a) \mathbb{R} is not a lc-regular space. To see this let A=[0,1)A=[0,1). Then AA is locally closed and 1A1\notin A. But AA cannot be separated from 11 by disjoint open sets.
b) \mathbb{R} is not a lc-completely regular space. To see this let A={1,12,}A=\{1,\frac{1}{2},\cdots\}. Then AA is locally closed and 0A0\notin A. But AA cannot be separated from 0 by a continuous function.
c) \mathbb{R} is not a lc-normal space. To see this let A={1,12,}A=\{1,\frac{1}{2},\cdots\} and B={0}B=\{0\}. Then AA and BB are locally closed sets. But AA and BB cannot be separated by disjoint open sets.

Remark 4.3.

a) Every lc-completely regular TDT_{D}-space is completely regular.
b) Every lc-normal TDT_{D}-space is Hausdorff.

Definition 4.4.

A space XX is called lc-compact if each locally closed cover of XX has a finite subcover.

Every lc-compact space is compact. Every infinite compact T1T_{1}-space is not lc-compact. One can easily see that (X,𝒯)(X,{\mathcal{T}}) is lc-compact if and only if (X,𝒯l)(X,{\mathcal{T}_{l}}) is compact.

Proposition 4.5.

Every clopen subset of a lc-compact space is lc-compact.

Proof.

Suppose that YαΛAαΛY\subseteq\bigcup_{\alpha\in\Lambda}A_{\alpha}\in\Lambda, where AαA_{\alpha} is locally closed set in YY, for each αΛ\alpha\in\Lambda. Hence, Aα=BαYA_{\alpha}=B_{\alpha}\cap Y, which BαB_{\alpha} is locally closed set in XX, for each αΛ\alpha\in\Lambda. Therefor there is an open set GαG_{\alpha} and a closed set FαF_{\alpha} in XX which Bα=GαFαB_{\alpha}=G_{\alpha}\cap F_{\alpha}. Now it is clear that X=αΛ((Gα(XY))(Fα(XY)))X=\bigcup_{\alpha\in\Lambda}((G_{\alpha}\cup(X-Y))\cap(F_{\alpha}\cup(X-Y))). Since XX is a lc-compact space we infer that X=k=1n((Gαk(XY))(Fαk(XY)))X=\bigcup_{k=1}^{n}((G_{\alpha_{k}}\cup(X-Y))\cap(F_{\alpha_{k}}\cup(X-Y))), for a natural number nn and αiΛ\alpha_{i}\in\Lambda for i=1,,ni=1,\cdots,n . It implies that Yk=1nAαkY\subseteq\bigcup_{k=1}^{n}A_{\alpha_{k}}, that is, YY is a lc-compact space. ∎

Definition 4.6.

A function f:XYf:X\to Y is called lc-continuous if the converse image of any open set in YY is locally closed in XX.

Every continuous function is lc-continuous. The converse is not true. For example, let X={a,b}X=\{a,b\} and 𝒯1={,{a},X}{\mathcal{T}_{1}}=\{\emptyset,\{a\},X\} and 𝒯2={,{a},{b},X}{\mathcal{T}_{2}}=\{\emptyset,\{a\},\{b\},X\} are two topology on XX. We define f:(X,𝒯1)(X,𝒯2)f:(X,{\mathcal{T}_{1}})\to(X,{\mathcal{T}_{2}}) by f(x)=xf(x)=x. Then ff is lc-continuous but it is not a continuous function.

Remark 4.7.

Let f:XYf:X\to Y be a onto and XX be a lc-compact space. Then
a) if ff is continuous then YY is lc-compact.
b) if ff is lc-continuous then YY is compact.

Remark 4.8.

If f:(X,𝒯)Yf:(X,{\mathcal{T}})\to Y is lc-continuous, then f:(X,𝒯l)Yf:(X,{\mathcal{T}_{l}})\to Y is continuous. The converse is false. To see this we consider (,𝒯u)(\mathbb{R},{\mathcal{T}_{u}}) and define f:(,𝒯l)Y={0,1}f:(\mathbb{R},{\mathcal{T}_{l}})\to Y=\{0,1\} by f()=0f(\mathbb{Q})=0 and f()=1f(\mathbb{R}-\mathbb{Q})=1, which YY is equipped with discrete topology. Since (,𝒯l)(\mathbb{R},{\mathcal{T}_{l}}) is discrete, then ff is continuous. But =f1({0})\mathbb{Q}=f^{-1}(\{0\}) is not locally closed set in (,𝒯u)(\mathbb{R},{\mathcal{T}_{u}}) while {0}\{0\} is an open set in YY. This shows that f:(,𝒯u)Yf:(\mathbb{R},{\mathcal{T}_{u}})\to Y is not a lc-continuous function.

Remark 4.9.

Let f:XYf:X\to Y be continuous.Then
a) if BYB\subseteq Y is locally closed then f1(B)Xf^{-1}(B)\subseteq X is locally closed.
b) if AXA\subseteq X is locally closed then f(A)f(A) need not be locally closed. Let X={a,b}X=\{a,b\}, 𝒯1{\mathcal{T}_{1}} be discrete topology and 𝒯2{\mathcal{T}_{2}} be trivial topology on XX. We define f:(X,𝒯1)(X,𝒯2)f:(X,{\mathcal{T}_{1}})\to(X,{\mathcal{T}_{2}}) by f(x)=xf(x)=x. Then A={a}A=\{a\} is locally closed set in (X,𝒯1)(X,{\mathcal{T}_{1}}) but f(A)={a}f(A)=\{a\} is not locally closed set in (X,𝒯2)(X,{\mathcal{T}_{2}}).

Definition 4.10.

A function f:XYf:X\to Y is called locally closed if the image of each locally closed set of XX, is locally closed in YY.

If f:XYf:X\to Y is one-to-one, open and closed function, then it is locally closed. A locally closed function need not be open or closed. See the next example.

Example 4.11.

a) Let X={a,b}X=\{a,b\}, 𝒯1{\mathcal{T}_{1}} be discrete topology and 𝒯2={,{a},X}{\mathcal{T}_{2}}=\{\emptyset,\{a\},X\} be another topology on XX. We define f:(X,𝒯1)(X,𝒯2)f:(X,{\mathcal{T}_{1}})\to(X,{\mathcal{T}_{2}}) by f(x)=xf(x)=x. Then ff is locally closed but it is not an open function.
b) Let X={a,b,c}X=\{a,b,c\}, 𝒯1{\mathcal{T}_{1}} be discrete topology and 𝒯2={,{a,c},{b,c},{c},X}{\mathcal{T}_{2}}=\{\emptyset,\{a,c\},\{b,c\},\{c\},X\} be another topology on XX. We define f:(X,𝒯1)(X,𝒯2)f:(X,{\mathcal{T}_{1}})\to(X,{\mathcal{T}_{2}}) by f(x)=xf(x)=x. Then ff is locally closed but it is not a closed function.

References

  • [1] A. R. Aliabad, V. Bagheri and M. Karavan Jahromi, On Quasi PP-Spaces and Their Application in Submaximal and Nodec Spaces, Bull. Iranian Math. Soc. Vol. 43, No. 3, (2017), 835-852.
  • [2] A. V. Arhangel’skii and P. J. Collins, On submaximal spaces, Top. Appl. 64 (1995), No. 3, 219-241.
  • [3] C. E. Aull and W. J. Thron, Separation axioms between T0T_{0} and T1T_{1} , Indagationes Math. 24 (1962), 26–37.
  • [4] F. Azarpanah, Intersection of essential ideals in C(X)C(X), Proc. Amer. Math. Soc. 125 (1997), 2149-2154.
  • [5] F. Azarpanah and T. Soundararajan, When the family of functions vanishing at infinity is an ideal of C(X)C(X), Rocky Mountain J. Math., Vol. 31, No. 4 (2001), 1133-1140.
  • [6] F. Azarpanah, R. Mohamadian, P. Monjezi, On PFP_{F}-spaces, Top. Appl. 302 (2021), 1-8.
  • [7] K. Belaid, L. Dridi, O. Echi, Submaximal and door compactifications, Top. Appl., 158 (2011), 1969-1975.
  • [8] J. Dontchev, On submaximal spaces, Tamkang J. Math., Vol. 26, No 3, (1995).
  • [9] L. Dridi, S. Lazaar and T. Turki, FF-door spaces and FF-submaximal spaces, Applied General Topology, Volume 14, No. 1, (2013), 97-113
  • [10] R. Engelking, General Topology, PWN Polish Scientific Publishers, (1977).
  • [11] M. Ganster and I. L. Reilly, locally closed sets and LC-continuous functions, Internat. J. Math. Sci. vol 12, No 3, (1989), 417-424.
  • [12] L. Gillman, M. Jerison, Rings of Continuous Functions, Springer-Verlag, 1976.
  • [13] D. Johnson and M. Mandelker, Functions with pseudocompact support, Gen. Topology Appl. 3 (1973), 331-338.
  • [14] O. A. S. Karamzadeh and M. Rostami, On the Intrinsic Topology and Some Related Ideals of C(X), Proc. Amer. Math. Soc., Vol. 93, No. 1, (1985), 179-184.
  • [15] F. Manshoor, Characterization of some kind of compactness via some properties of the space of functions with mm-topology, Int. J. Contemp. Math. Sciences, 7, No. 21 (2012), 1037-1042.
  • [16] M. Rajagopalan and R. F. Wheeler, Sequential compactness of X implies a compactness property for C(X), Canad J. Math. 28 (1976), 207-210.
  • [17] B. Richmond, Principal topologies and transformation semigroups, Top. Appl. 155 (2008) 1644–1649.
  • [18] R. C. Walker, The Stone-Čech Compactification, Springer, New York, 1974.
  • [19] S. Willard, General Topology, Addison- Wesley, 1970.