@ Appl. Gen. Topol. 16, no. 2(2015), 183-207doi:10.4995/agt.2015.3445 c© AGT, UPV, 2015 On the locally functionally countable sub algebra of C(X) O. A. S. Karamzadeh, M. Namdari and S. Soltanpour Department of Mathematics, Shahid Chamran University of Ahvaz, Ahvaz, Iran (karamzadeh@ipm.ir, namdari@ipm.ir, s-soltanpour@phdstu.scu.ac.ir) Abstract Let Cc(X) = {f ∈ C(X) : |f(X)| ≤ ℵ0}, C F (X) = {f ∈ C(X) : |f(X)| < ∞}, and Lc(X) = {f ∈ C(X) : Cf = X}, where Cf is the union of all open subsets U ⊆ X such that |f(U)| ≤ ℵ0, and CF (X) be the socle of C(X) (i.e., the sum of minimal ideals of C(X)). It is shown that if X is a locally compact space, then Lc(X) = C(X) if and only if X is locally scattered. We observe that Lc(X) enjoys most of the important properties which are shared by C(X) and Cc(X). Spaces X such that Lc(X) is regular (von Neumann) are characterized. Similarly to C(X) and Cc(X), it is shown that Lc(X) is a regular ring if and only if it is ℵ0-selfinjective. We also determine spaces X such that Soc ( Lc(X) ) = CF (X) (resp., Soc ( Lc(X) ) = Soc ( Cc(X) ) ). It is proved that if CF (X) is a maximal ideal in Lc(X), then Cc(X) = C F (X) = Lc(X) ∼= n ∏ i=1 Ri, where Ri = R for each i, and X has a unique infinite clopen connected subset. The converse of the latter result is also given. The spaces X for which CF (X) is a prime ideal in Lc(X) are characterized and consequently for these spaces, we infer that Lc(X) can not be isomorphic to any C(Y ). 2010 MSC: Primary: 54C30; 54C40; 54C05; 54G12; Secondary: 13C11; 16H20. Keywords: functionally countable space; socle, zero-dimensional space; scattered space; locally scattered space, ℵ0-selfinjective. Received 9 December 2014 – Accepted 22 April 2015 http://dx.doi.org/10.4995/agt.2015.3445 O. A. S. Karamzadeh, M. Namdari and S. Soltanpour 1. Introduction C(X) denotes the ring of all real valued continuous functions on a topo- logical space X. In [10] and [11], Cc(X), the subalgebra of C(X), consisting of functions with countable image are introduced and studied. It turns out that Cc(X), although not isomorphic to any C(Y ) in general, enjoys most of the important properties of C(X). This subalgebra has recently received some attention, see [10], [23], [24], [4], and [11]. Since Cc(X) is the largest subring of C(X) whose elements have countable image, this motivates us to consider a natural subring of C(X), namely Lc(X), which lies between Cc(X) and C(X). Our aim in this article, similarly to the main objective of working in the con- text of C(X), is to investigate the relations between topological properties of X and the algebraic properties of Lc(X). In particular, we are interested in finding topological spaces X for which Lc(X) = C(X). An outline of this pa- per is as follows: In Section 2, we show that if X is a locally compact space, then Lc(X) = C(X) if and only if X is locally scattered, which is somewhat similar to a classical result due to Rudin in [27], and Pelczynski and Semadeni in [25] (of course, by no means as significant). This classical result says that a compact space X is scattered if and only if C(X) = Cc(X). Let us for the sake of the brevity, call the latter classical result, RPS-Theorem. If X is an almost discrete space or a P-space, then L1(X) = LF (X) = Lc(X) = C(X), where LF (X) and L1(X) are the locally functionally finite (resp., constant) subalgebra of C(X), see Definition 2.7. In Section 3, we introduce zl-ideals in Lc(X) and trivially observe that most of the facts related to z-ideals are extendable to zl-ideals. In Section 4, topolog- ical spaces in which points and closed sets are separated by elements of Lc(X), are called locally countable completely regular space (briefly, lc-completely reg- ular). Clearly, every zero-dimensional space is lc-completely regular (note, in the zero-dimensional case, points and closed sets are separated even by the ele- ments of Cc(X), which is a subring of Lc(X)), see [10, Proposition 4.4]. Spaces X, for which Lc(X) is regular, are called locally countably P-space (briefly, LCP-space) and are characterized both algebraically and topologically in this section. It is shown that P-spaces and LCP-spaces coincide when X is lc- completely regular. Finally, in this section similar to C(X) and Cc(X), we prove that Lc(X) is a regular ring if and only if it is ℵ0-selfinjective. The socle of C(X) (i.e., CF (X)) which is in fact a direct sum of minimal ideals of C(X) is characterized topologically in [20, Proposition 3.3], and it turns out that CF (X) is a useful object in the context of C(X), see [20], [1], [2], [8], [3], and [6]. The socle of Cc(X), denoted by Soc ( Cc(X) ) , is studied in [11, Propo- sition 5.3], and spaces X for which Soc ( Cc(X) ) = CF (X) are determined in [11, Theorem 5.6]. Motivated by the latter facts, we characterize the socle of Lc(X) both topologically and algebraically, in Section 5. Spaces X for which Soc ( Lc(X) ) = Soc ( Cc(X) ) and Soc ( Lc(X) ) = CF (X) are also characterized. In [8, Proposition 1.2], [3, Remark 2.4], it is shown that CF (X) can not be a prime ideal in C(X), where X is any space. But, in [11, Proposition 6.2], spaces c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 184 On the locally functionally countable sub algebra of C(X) X such that CF (X) is prime in Cc(X) are characterized. The latter character- ization is similarly extended to Lc(X). Consequently, this implies that Lc(X) is not isomorphic to any C(Y ) in general. All topological spaces that appear in this article are assumed to be infinite completely regular Hausdorff, unless otherwise mentioned. For undefined terms and notations the reader is referred to [13], [7]. 2. The subalgebra Lc(X) of C(X) Definition 2.1. Let f ∈ C(X) and Cf be the union of all open sets U ⊆ X, such that f(U) is countable. We define Lc(X) to be the set of all f ∈ C(X) such that Cf is dense in X, i.e., Cf = ⋃ {U| U is open in X and |f(U)| ≤ ℵ0} Lc(X) = {f ∈ C(X) : Cf = X} We shall briefly and easily notice that, Lc(X) is a subalgebra as well as a sublattice of C(X) containing Cc(X), and we call it the locally functionally countable subalgebra of C(X). It is manifest that CF (X) ⊆ CF (X) ⊆ Cc(X) ⊆ Lc(X) ⊆ C(X), where CF (X) = {f ∈ C(X) : |f(X)| < ∞}, see [10]. The following example shows that the equality between any two of these objects may not necessarily hold. Example 2.2. Let the basic neighborhood of x be the set {x}, for each point x ≥ √ 2 and for the rest of the real numbers (i.e., x < √ 2) the basic neigh- borhoods be the usual open intervals containing x. This is a topology T on R and in this case we put X = R. Clearly, X is a completely regular Hausdorff space which is finer than the usual topology of R. The function f : X → R, where f(x) = 1 for x ≥ √ 2, and f(x) = 0 otherwise, is continuous and X\Z(f) is infinite, hence f ∈ CF (X)\CF (X), see [20, Proposition 3.3]. We define g : X → R, such that g(x) = x for x ∈ [ √ 2, ∞) ∩ Q and g(x) = 0 for x ∈ ([ √ 2, ∞) ∩ Qc) ∪ (−∞, √ 2), hence g ∈ Cc(X)\CF (X). Also we observe that for the function h : X → R, where h(x) = x for x ≥ √ 2, and h(x) = √ 2 otherwise, we have h ∈ Lc(X)\Cc(X). The identity function i : X → R is continuous and Ci = [ √ 2, ∞), see Definition 2.1. Hence i ∈ C(X)\Lc(X). We note that Cf = X if and only if for every open subset G ⊆ X, there exists an open subset U ⊆ X such that |f(U)| ≤ ℵ0 and U ∩ G 6= ∅ or equivalently if and only if for each open subset G ⊆ X, there exists a nonempty open subset V ⊆ G with |f(V )| ≤ ℵ0. Lemma 2.3. For the space X the following statements hold. (1) If f, g ∈ C(X), then Cf+g ⊇ Cf ∩ Cg. (2) If f, g ∈ C(X), then Cfg ⊇ Cf ∩ Cg. (3) If f ∈ C(X), then C|f| = Cf . (4) If f ∈ C(X), then C 1 f = Cf . (5) If f, g ∈ Lc(X), then Cf ∩ Cg = X. c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 185 O. A. S. Karamzadeh, M. Namdari and S. Soltanpour Proof. Let Cf = ⋃ U⊆X |f(U)|≤ℵ0 U and Cg = ⋃ V ⊆X |g(V )|≤ℵ0 V , where U and V are open subsets of X, then Cf ∩ Cg = ⋃ U,V ⊆X |f(U)|,|g(V )|≤ℵ0 (U ∩ V ) Hence (1), (2), (3), (4) are evident. For part (5) we recall that if Y is a dense subset of X and G is an open subset of X, then G ∩ Y = G. Since Cf , Cg are open and dense in X we infer that Cf ∩ Cg = Cf = Cg = X. � The following examples show that the equalities in (1), (2) of the previous lemma do not necessarily hold, in general. Example 2.4. (1) Let i : R → R be the identity function and f : R → R with f(x) = −x, then Ci = Cf = ∅, but Ci+f = R. Hence Ci+f ) Ci ∩ Cf . (2) Let i : R\{0} → R be the identity function and f : R\{0} → R with f(x) = 1/x, then Ci = Cf = ∅, but Cif = R\{0}. Hence Cif ) Ci ∩ Cf . The following fact shows that Lc(X) is indeed a subalgebra of C(X) such that whenever Z(f) = ∅ where f ∈ Lc(X), then f is a unit in Lc(X). We remind the reader that the latter fact is not true for C∗(X). Corollary 2.5. For the space X the following statements hold. (1) If f, g ∈ Lc(X), then f + g ∈ Lc(X) and fg ∈ Lc(X). (2) f ∈ Lc(X) if and only if |f| ∈ Lc(X). (3) Let f be a unit element in C(X), then f ∈ Lc(X) if and only if 1f ∈ Lc(X). Corollary 2.6. Lc(X) is a sublattice of C(X). Definition 2.7. Let f ∈ C(X) and CFf be the union of all open sets U ⊆ X such that f(U) is finite. We define LF (X) to be the set of all f ∈ C(X) such that CFf is dense in X, and call it locally functionally finite subalgebra of C(X), i.e., CFf = ⋃ {U| U is open in X and |f(U)| < ∞} LF (X) = {f ∈ C(X) : CFf = X} In particular, let f ∈ C(X) and Ccf be the union of all open sets U ⊆ X such that f(U) is constant. We define L1(X) to be the set of all f ∈ C(X) such that Ccf is dense in X, and we call it locally functionally constant subalgebra of C(X), i.e., Ccf = ⋃ {U| U is open in X and |f(U)| = 1} L1(X) = {f ∈ C(X) : Ccf = X} Clearly, LF (X) and L1(X) are subalgebras of Lc(X). In [26] and [15], E0(X) is defined, and by the above notation we have E0(X) = L1(X). It is evident that CF (X) ⊆ LF (X). c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 186 On the locally functionally countable sub algebra of C(X) Remark 2.8. We note that Lemma 2.3, Corollary 2.5, and Corollary 2.6 are also valid for LF (X) and L1(X). Remark 2.9. It is manifest that Cc(X) = R, where X = [0, 1]. But the Can- tor function f is a monotonic nonconstant continuous function, and Ccf = [0, 1]\C = [0, 1], where C is the Cantor set, see [9], [5]. Therefore the Cantor function f belongs to L1([0, 1]), and R ( L1([0, 1]), hence R ( Lc([0, 1]). We emphasize that Cc(X) = R, but R ( Lc(X), and this can be considered as an advantage of Lc(X) over Cc(X), in this case. Remark 2.10. In [15], a first countable compact space X (resp., in [26], a non- first countable compact space X) is constructed such that L1(X) = R. We are interested in characterizing topological spaces X for which Lc(X) = C(X). In the following proposition we have a simple result, which is similar to RPS-Theorem. Let us recall that in a commutative ring R by an annihilator ideal I, we mean I = Ann(S) = {r ∈ R : rS = 0}, where S 6= {0} is a nonempty subset of R. Proposition 2.11. If X is an almost discrete space (i.e., I(X), the set of isolated points of X, is dense in X), then L1(X) = LF (X) = Lc(X) = C(X). In particular, if every nonzero annihilator ideal of C(X), where X is any space, contains a nonzero minimal ideal, then the latter equalities hold. Proof. If f ∈ C(X), then Ccf ⊇ ⋃ x∈I(X){x} = I(X). Hence Ccf = X, i.e., f ∈ L1(X). Finally, we first recall that C(X) contains many nonzero zero- divisors (note, for each 0 6= f ∈ C(X), (f − |f|)(f + |f|) = 0. Hence nontrivial annihilator ideals in C(X) always exist. Consequently, by our assumption the socle of C(X) is not zero, i.e., CF (X) 6= 0. We now claim that Ann ( CF (X) ) = 0. To see this, if I = Ann ( CF (X) ) 6= 0, then I must contain a nonzero minimal ideal, hence I ∩CF (X) 6= 0. But, (I ∩CF (X))2 = 0 and since C(X) is reduced, we infer that I ∩CF (X) = 0, which is absurd. This means that we have already shown that Ann ( CF (X) ) = 0, which by [20, Proposition 2.1] is equivalent to the density of I(X) in X, hence we are done by the first part. � Before, presenting the next fact, we evidently note that every scattered space is an almost discrete space, for if x ∈ X and Ux is a neighborhood of x, then Ux has an isolated point x0. Since Ux is open, x0 is an isolated point of X, too. Hence x0 ∈ Ux ∩ I(X) 6= ∅, therefore I(X) = X. Proposition 2.12. If X is a scattered space, then L1(X) = LF (X) = Lc(X) = C(X). In particular, if X is a compact scattered space, then the latter rings coincide with Cc(X). Proof. By the above comment and RPS-Theorem we are done. � The following example shows that the converse of the above corollary is not valid. c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 187 O. A. S. Karamzadeh, M. Namdari and S. Soltanpour Example 2.13. Let for each point x ∈ Q, the basic neighborhood of x be the singleton {x}, and for each x ∈ Qc, the basic neighborhood of x be the usual open interval containing x. This constitutes a topology on X = R, and it is clearly, a Hausdorff normal space which is almost discrete, since I(X) = Q. Hence Lc(X) = C(X), but X is not scattered. In view of RPS-Theorem we may naturally define a compact space X to be scattered if given any f ∈ C(X) and any x ∈ X, there exists a compact neighborhood Vf of x such that |f(V ◦f )| ≤ ℵ0. Motivated by this we give the following definition. Definition 2.14. A space X is called locally scattered if given any f ∈ C(X) and a nonempty open set G, there exists a compact subset Vf of X in G, with ∅ 6= V ◦f ⊆ G and |f(V ◦f )| ≤ ℵ0. The space βX where X is discrete is locally scattered. Clearly, every scat- tered space is a locally scattered space, but the converse is not true. For example, βN is a locally scattered space which is not scattered, for βN\N has no isolated point (note, each clopen subset of βN\N has the same cardinality as βN\N, see [13, 6S(4)]). Lemma 2.15. Let X be a locally scattered space. Then every open C-embedded subset of X (e.g., any clopen subset) is also locally scattered. Proof. Let Y be an open C-embedded subset of X, and G be an open subset in Y , and f ∈ C(Y ). Since Y is C-embedded in X, we infer that there exists g ∈ C(X) such that g|Y = f. Clearly, G is open in X and by our assumption, there exists a compact subset Vg in G such that ∅ 6= V ◦g ⊆ G ⊆ Y , |g(V ◦g )| ≤ ℵ0. Thus Vg is compact in Y in G with |f(V ◦g )| = |g(V ◦g )| ≤ ℵ0, i.e., Y is locally scattered. � Let us recall that a Hausdorff space X is locally compact if and only if each point in X has a compact neighborhood. Clearly, every compact Hausdorff space is locally compact. The following result is somewhat similar to RPS- Theorem. Theorem 2.16. Let X be a compact space. Then Lc(X) = C(X) if and only if X is locally scattered. In particular, if X is a discrete space and Y is a non-scattered clopen subset of βX (e.g., X = N and Y = βN), then Lc(Y ) = C(Y ) = C ∗(Y ) 6= Cc(Y ). Proof. First, we assume that X is compact and Lc(X) = C(X). Now, for each f ∈ C(X) we have Cf = X. Hence for any nonempty open subset G in X there exists an open subset Uf in X such that |f(Uf )| ≤ ℵ0, Uf ∩ G 6= ∅. Since the open subsets of a locally compact space are locally compact, we infer that Uf ∩ G is locally compact. Consequently, any neighborhood of a point x ∈ Uf ∩ G contains a compact neighborhood, Vf say, of x. Hence x ∈ V ◦f ⊆ Vf ⊆ Uf ∩ G ⊆ X and |f(V ◦f )| ≤ |f(Uf )| ≤ ℵ0, which means that X is locally scattered and we are done. The converse is evident by Definition 2.14, c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 188 On the locally functionally countable sub algebra of C(X) and the definition of Lc(X). For the last part, we notice that Y as a closed subset of βX is compact and by Lemma 2.15, it is locally scattered. Now by the first part and the compactness of Y , we have Lc(Y ) = C(Y ) = C ∗(Y ). But, in view of RPS-Theorem and the fact that Y is not scattered, we infer that Cc(Y ) 6= C(Y ), and we are done. � The previous proof immediately yields the following fact, too. Corollary 2.17. Let X be a locally compact space. Then Lc(X) = C(X) if and only if X is locally scattered. We recall that if the set of open neighborhoods of a point P in X is closed under countable intersection, then P is called a P-point. The set of all P-points of X is denoted by PX and X is called a P-space if PX = X. An interesting result due to A. W. Hager asserts that a P-space X is functionally countable (i.e., C(X) = Cc(X)) if and only if it is pseudo-ℵ1-compact (i.e., each locally finite family of open sets is countable), see [21, Proposition 3.2]. This result is extended to Cc(X) = C F (X) in [11, Proposition 4.2]. The following is also a counterpart of the latter result. Proposition 2.18. If PX = X (in particular, if X is a P-space), then L1(X) = LF (X) = Lc(X) = C(X). Proof. For each f ∈ C(X) and x ∈ PX there exists an open neighborhood Ux of x such that f is constant on Ux, see [13, 4L(3)]. Therefore C c f ⊇ ⋃ |f(Ux)|=1 Ux ⊇ PX, hence f ∈ L1(X). � We note that βN is not a P-space while L1(βN) = LF (βN) = Lc(βN) = C(βN). By [13, 6V(6)], βN\N has a dense set of P-points, hence L1(βN\N) = LF (βN\N) = Lc(βN\N) = C(βN\N). Remark 2.19. Let X be a P-space without isolated points, see [13, 13 P], then X is not almost discrete. But by Proposition 2.18, L1(X) = LF (X) = Lc(X) = C(X), see also Proposition 2.11. Let us borrow the following definition from [16]. Definition 2.20. A topological space X is called locally functionally countable if every point x ∈ X is countably P-point, in the sense that there exists an open neighborhood Ux of x such that C(Ux) = Cc(Ux). The following result implies that if a space X is second countable or a compact space, then X is locally functionally countable if and only if it is functionally countable (i.e., C(X) = Cc(X)). Proposition 2.21. Let X be a Lindelöf space. Then X is locally functionally countable if and only if it is functionally countable. Proof. It is evident that every functionally countable space is locally function- ally countable (note, for each x ∈ X take Ux = X). Conversely, let X be locally functionally countable, then for each f ∈ C(X), f(X) = f( ⋃ x∈X Ux), where c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 189 O. A. S. Karamzadeh, M. Namdari and S. Soltanpour Ux is an open neighborhood of x with C(Ux) = Cc(Ux). Since X is Lindelöf and C(X) ⊆ C(Ux), for each x ∈ X, we infer that f(X) = f( ⋃∞ i=1 Uxi) =⋃∞ i=1 f(Uxi) is countable, and we are done. � The next result shows that for every locally functionally countable space X, C(X) coincides with Lc(X). But the converse is not true in general, see Exam- ple 2.13 (note, R with the topology in this example is not locally functionally countable, for no irrational number is a countably P-point). Proposition 2.22. If X is a locally functionally countable space, then Lc(X) = C(X). Proof. We must show that for each f ∈ C(X), Cf = X. Let G ⊆ X be an open set in X and x ∈ G. Since X is locally functionally countable, there exists an open neighborhood Ux of x such that C(Ux) = Cc(Ux). Clearly |f(Ux)| = |(f|Ux)(Ux)| ≤ ℵ0. Now, x ∈ Ux ⋂ G 6= ∅ and Ux ⊆ Cf imply that Cf ⋂ G 6= ∅, hence Cf = X. � It is clear that if Y is a subset of X such that for each f ∈ C(X), f|Y is constant, then Y must be a singleton. For otherwise, if y1, y2 ∈ Y and y1 6= y2, then by complete regularity of X there exists f ∈ C(X) such that f(y1) 6= f(y2), which is absurd. Hence the following definition, which is also needed, is now in order. Definition 2.23. If Y is a subset of a space X, then the set of all f ∈ C(X) such that f|Y is constant is a subalgebra of C(X), denoted by C1(Y ). Naturally, we say that Y is constant with respect to a subring A of C(X) if A ⊆ C1(Y ). We note that for every topological space X, C1(X) = C(X) if and only if X is singleton. If Y is a proper closed subset of X, then R ( C1(Y ). The following proposition is evident. Proposition 2.24. Let X be a topological space and Y be a connected subset of X, then Cc(X) ⊆ C1(Y ). In particular, if X\Y is countable, then A ⊆ C1(Y ) if and only if A ⊆ Cc(X). We conclude this section with the following fact whose proof is evident by the complete regularity of X. Corollary 2.25. For any subspace Y of X, R ⊆ C1(Y ) ⊆ C(X). Moreover, C1(Y ) = R if and only if Y is dense in X. Proof. For the last part we note that if x /∈ Y , then there exists f ∈ C(X) with f(x) = 0 and f(Y ) = 1, i.e., C1(Y ) 6= R. This implies that Y = X in case C1(Y ) = R. Conversely, let Y = X and take f ∈ C(X) such that f ∈ C1(Y ), then f(Y ) = c, where c ∈ R. Consequently, f = c in C(X), for Y is dense in X, hence we are done. � c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 190 On the locally functionally countable sub algebra of C(X) 3. zl-IDEALS We remind the reader that many facts in the context of C(X) can be ex- tended naturally to Lc(X), similarly to Cc(X), see [10]. The proofs of most of the results in this section follow mutatis mutandis from the proofs of their corresponding results in [10]. Therefore, we state them without proofs, for the record, but give pertinent references for their corresponding proofs (note, the reason that we emphasize on the recording of these facts here is because we do believe that Lc(X) and Cc(X), are eligible to play appropriate roles as com- panions of C(X), in the future studies in the context of C(X), see for example, the comment in the first two lines of the introduction in [4]. Definition 3.1. A space X is said to be locally countably pseudocompact (briefly, lc-pseudocompact) if L∗c(X) = Lc(X), where L ∗ c(X) = Lc(X)∩C∗(X). The next three results are the counterparts of [13, Theorem 1.7, Corollary 1.8, and Theorem 1.9]. Proposition 3.2. Every homomorphism ϕ : Lc(X) → Lc(Y ) takes L∗c(X) into L∗c(Y ). Corollary 3.3. If Y is not a lc-pseudocompact space, then Lc(Y ) can not be a homomorphic image of any L∗c(X). Corollary 3.4. Let ϕ be a homomorphism from Lc(X) into Lc(Y ) whose image contains L∗c(Y ), then ϕ(L ∗ c(X)) = L ∗ c(Y ). If f ∈ Lc(X) and f > 0, then there exists g ∈ Lc(X) with f = g2. We also note that whenever f ∈ Lc(X) and fr ∈ C(X) where r ∈ R, then fr ∈ Lc(X). We recall that all positive units in Lc(X) have the same number of square roots, see [13, 1B(1)]. The following proposition and its corollary are the counterparts of [13, 1D(1)] and [10, Lemma 2.4] for Lc(X). Since the latter facts play a basic role in the context of C(X), we present sketch of proofs for these counterparts. Proposition 3.5. If f, g ∈ Lc(X) and Z(f) is a neighborhood of Z(g), then f = gh for some h ∈ Lc(X). Proof. We have Zl(g) ⊆ intZl(f). Put h(x) = { 0 , x ∈ Zl(f) f(x) g(x) , x /∈ intZl(f) therefore h ∈ C(X), and Ch ⊇ Cf ∩ C1/g = Cf ∩ Cg = X. Hence h ∈ Lc(X) and f = gh. � Corollary 3.6. If f, g ∈ Lc(X), and |f| ≤ |g|r, r > 1, then f = gh for some h ∈ Lc(X). In particular, if |f| ≤ |g|, then whenever fr is defined for r > 1, fr is a multiple of g. Proof. Let h(x) = { 0 , x ∈ Zl(g) f(x) g(x) , x /∈ Zl(g) c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 191 O. A. S. Karamzadeh, M. Namdari and S. Soltanpour then h ∈ C(X), Ch ⊇ Cf ∩ Cg = X. Hence h ∈ Lc(X) and f = gh. � Proposition 3.7. If f ∈ Lc(X), then there exists a positive unit u ∈ Lc(X) with (−1 ∨ f) ∧ 1 = uf. Proof. Put u(x) = { 1 , − 1 ≤ f(x) ≤ 1 1 |f(x)| , 1 ≤ |f(x)| Clearly Cu = Cf , hence u ∈ Lc(X) and (−1 ∨ f) ∧ 1 = uf. So if f ∈ Lc(X), then f and (−1 ∨ f) ∧ 1 belongs to an ideal of Lc(X) � Remark 3.8. The previous results are also true if we replace Lc(X) by either LF (X) or L1(X). Convention. Let us put Zl(X) = {Z(f) : f ∈ Lc(X)}, ZF (X) = {Z(f) : f ∈ LF (X)}, and Z1(X) = {Z(f) : f ∈ L1(X)}, where X is a topological space. Definition 3.9. Two subsets A and B of a topological space X are said to be locally countably separated (briefly, lc-separated) in X if there is an element f ∈ Lc(X) such that f(A) = 1, f(B) = 0. The following result is the counterpart of [13, Theorem 1.15], [10, Theorem 2.8] . Theorem 3.10. Two subsets A, B of a space X are lc-separated if and only if they are contained in disjoint members of Zl(X). Moreover, lc-separated sets have disjoint zero-set neighborhoods in Zl(X). Clearly, if a < b and f ∈ Lc(X) such that f(x) ≤ a, ∀x ∈ A, and f(x) ≥ b, ∀x ∈ B, where A, B are subsets of X, then A, B are lc-separated in X. Corollary 3.11. If A, B are lc-separated in X, then there are zero-sets Z1, Z2 in Zl(X) with A ⊆ X \ Z1 ⊆ Z2 ⊆ X \ B. Definition 3.12. ∅ 6= F ⊆ Zl(X) is called a zl-filter on X if F satisfies the following conditions. (1) ∅ /∈ F . (2) Z1, Z2 ∈ F , then Z1 ∩ Z2 ∈ F . (3) Z ∈ F , Z′ ∈ Zl(X) with Z′ ⊇ Z, then Z′ ∈ F . Prime zl-filter and zl-ultrafilter are defined similarly to their counterparts in [13]. If I is an ideal of Lc(X), then Zl[I] = {Z(f) : f ∈ I} is a zl-filter on X. Conversely, if F is a zl-filter on X, then Z −1[F ] = {f ∈ Lc(X) : Z(f) ∈ F} is an ideal in Lc(X). Moreover, every zl-filter F is of the form F = Zl[I] for some ideal I in Lc(X) and for any ideal J in Lc(X), Z −1[Zl[J]] is an ideal in Lc(X) containing J. In Example 2.13, we consider the identity function i : (R, T ) → R, clearly i ∈ Lc(R) = C(R). Now, put I = (i), then Zl(I) = {0}. Clearly, f(x) = x1/3 ∈ Lc(R), f ∈ Z−1[Zl[I]]\I. Hence the following definition is in order. c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 192 On the locally functionally countable sub algebra of C(X) Definition 3.13. An ideal I in Lc(X) is called a zl-ideal if whenever Z(f) ∈ Zl[I] and f ∈ Lc(X), then f ∈ I. Similarly, zF -ideal and z1-ideal are defined, see the previous convention. Clearly, every zl-ideal is an intersection of prime ideals in Lc(X). Similarly, every zF -ideal and z1-ideal is an intersection of prime ideals in LF (X) and L1(X). We emphasize again that the proofs of the following results are the same as the proofs of their counterparts in C(X) and Cc(X), see [10], and [13]. One can easily observe that these results up to Proposition 3.24 and including it are also valid for Lc(X) and LF (X). The next theorem is the counterpart of [13, Theorem 2.9], [10, Theorem 2.13]. Theorem 3.14. Let P be any zl-ideal in Lc(X). Then the following statements are equivalent. (1) P is a prime ideal in Lc(X). (2) P contains a prime ideal in Lc(X). (3) For all f, g ∈ Lc(X), if fg = 0, then f ∈ P or g ∈ P. (4) For each f ∈ Lc(X), there exists a zero-set in Zl[P ] on which f does not change sign. Corollary 3.15. Every prime ideal in Lc(X) is contained in a unique maximal ideal in Lc(X). Clearly if P is a prime ideal in Lc(X), then Zl[P ] is a prime zl-filter, and if F is a prime zl-filter, then Z −1 l [F ] is a prime zl-ideal. It is evident that every prime zl-filter is contained in a unique zl-ultrafilter. The following lemma is the counterpart of [10, Lemma 3.1], also see [28]. Lemma 3.16. Let f, g, l ∈ Lc(X), Z(f) ⊇ Z(g) ∩ Z(l) and define h(x) = { 0 , x ∈ Z(g) ∩ Z(l) fg2 g2+l2 , x /∈ Z(g) ∩ Z(l) , k(x) = { 0 , x ∈ Z(g) ∩ Z(l) fl2 g2+l2 , x /∈ Z(g) ∩ Z(l) Then we have the following conditions. (1) |k| ∨ |h| ≤ |f|. (2) f = h + k. (3) fl2 = k(g2 + l2), fg2 = h(g2 + l2). (4) h, k ∈ Lc(X). (5) Ch ⊇ Cf ∩ Cg ∩ Cl and Ck ⊇ Cf ∩ Cg ∩ Cl. The following results are the counterparts of [10, Corollary 3.2 to Corollary 3.8]. Lemma 3.17. Let A, B be two zl-ideals in Lc(X). Then either A+B = Lc(X) or A + B is a zl-ideal. Corollary 3.18. Let F = {Ai}i∈I be a collection of zl-ideals in Lc(X). Then either ∑ i∈I Ai = Lc(X) or ∑ i∈I Ai is a zl-ideal. c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 193 O. A. S. Karamzadeh, M. Namdari and S. Soltanpour Proposition 3.19. Every minimal prime ideal in Lc(X) is a zl-ideal. Corollary 3.20. Let F = {Pi}i∈I be a collection of minimal prime ideals in Lc(X). Then either ∑ i∈I Pi = Lc(X) or P = ∑ i∈I Pi is a prime ideal in Lc(X). Proposition 3.21. A prime ideal P in Lc(X) is absolutely convex. Proposition 3.22. The sum of a collection of semiprime ideals in Lc(X) is a semiprime ideal or is the entire ring Lc(X). Proposition 3.23. Let P be a prime ideal in Lc(X). Then the ring Lc(X)/P is totally ordered and its prime ideals are comparable. The next corollary is much stronger than Corollary 3.20 whose proof is similar to [10, Corollary 3.9]. Proposition 3.24. Let {Pi}i∈I be a collection of semiprime ideals in Lc(X) such that at least one of Pi’s is prime, then ∑ i∈I Pi is a prime ideal or all of Lc(X). All the previous results beginning with Theorem 3.14, are also valid for L1(X). The following theorem is the counter part of [10, Theorem 3.10], see also the comment preceding [10, Theorem 3.10]. Theorem 3.25. Let I be an ideal in Lc(X). Then I and √ I have the same largest zl-ideal. 4. LOCALLY COUNTABLE COMPLETELY REGULAR SPACES Definition 4.1. A Hausdorff space X is called locally countable completely regular (briefly, lc-completely regular) if whenever F ⊆ X is a closed set and x ∈ X\F , then there exists f ∈ Lc(X) with f(F) = 0 and f(x) = 1. We should remind the reader that, in this section, whenever the proof of a result is very similar to the proof of its counterpart in the literature, the proof is avoided. The proof of the following result is evident. Proposition 4.2. A Hausdorff space X is lc-completely regular if and only if whenever F ⊆ X is closed and x ∈ X \ F, then x and F have two disjoint zero-set neighborhoods in Zl(X). Consequently, there exist g, h ∈ Lc(X) with x ∈ X \ Z(h) ⊆ Z(g) ⊆ X \ F. Clearly X is a lc-completely regular space if and only if F = {Z(f) : f ∈ Lc(X)} is a base for the closed sets in X or equivalently if and only if B = {int(Z(f)) : f ∈ Lc(X)} is a base for the open sets in X. The next proposition is the counterpart of [13, 3.11(a)], [10, Proposition 4.3]. Proposition 4.3. Let X be a lc-completely regular space and A, B be two disjoint closed sets in X such that A is compact, then there is f ∈ Lc(X) with f(A) = 0 and f(B) = 1. c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 194 On the locally functionally countable sub algebra of C(X) Proposition 4.4. Let X be a compact space. Then X is lc-completely regular if and only if Lc(X) separates points in X. Since Cc(X) is a subring of Lc(X), the next result is evident, see [10, Propo- sition 4.4] Proposition 4.5. If X is a zero-dimensional space, then X is a lc-completely regular space. The following fact is also similar to [13, Theorem 3.6], and [10, Corollary 4.5]. Proposition 4.6. Let X be a Hausdorff space. Then X is a lc-completely regular space if and only if its topology coincides with the weak topology induced by Lc(X). We recall that X is a P-space (resp., CP-space) if and only if C(X) (resp., Cc(X)) is a regular ring, see [13, 4J] and [10]. In [10], it is shown that if C(X) is regular, then so too is Cc(X). If X is zero-dimensional, then the regularity of C(X) and Cc(X) coincide. We have already observed, see Proposition 2.18, that if X is a P-space, then L1(X) = LF (X) = Lc(X) = C(X). The next definition is now in order. Definition 4.7. A space X is called a locally countably P-space (briefly, LCP-space) if Lc(X) is regular. By the above comment we have the following result. Proposition 4.8. Every P-space is LCP-space. Proposition 4.9. If A is any regular subring of C(X) such that Cc(X) ⊆ A ⊆ C(X), then Cc(X) is regular. In particular, if Lc(X) is regular, then Cc(X) is regular, too. Proof. Let A be a regular ring, we must show that for each f ∈ Cc(X), there exists g ∈ Cc(X) such that f = f2g. Since A is regular, there is h ∈ A with f = f2h. Consequently, f = f2g, where g = h2f. It is also evident that Z(f) ⊆ Z(g) and g(x) = 1 f(x) , whenever x /∈ Z(f). Hence |g(X)| = |f(X)| , i.e., g ∈ Cc(X), and we are done. � Corollary 4.10. Let X be a zero-dimensional space. Then X is P-space if and only if any of the rings Cc(X), Lc(X) is regular. Remark 4.11. It is wroth mentioning that if X is a zero-dimensional space, then the regularity of C(X), Lc(X), Cc(X), and C(X, K) (where C(X, K), is a subring of C(X) whose elements take values in K, a subfield of R) coincide, see the above proposition and [10, Remark 7.5]. The following theorem is the counterpart of [10, Theorem 5.5] and its proof is also the same as the proof of its counterpart. We present a proof for the sake of completeness. c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 195 O. A. S. Karamzadeh, M. Namdari and S. Soltanpour Theorem 4.12. A space X is a LCP-space if and only if every zero-set in Zl(X) is open. Moreover, in this case whenever {fi}∞i=1 is a countable set in Lc(X), then ∞⋂ i=1 Zl(fi) is an open zero-set in Zl(X). Proof. Let X be a LCP-space and f ∈ Lc(X), hence f = f2g for some g ∈ Lc(X). It is evident that e = fg is an idempotent in Lc(X) and Z(f) = Z(e) = X\Z(1 − e) is clopen. Conversely, let Z(f) be open for each f ∈ Lc(X), we are to show that Lc(X) is regular. Since Z(f) = Z(f 2) for all f ∈ Lc(X), we infer that f = f2g for some g ∈ Lc(X), by Proposition 3.5. Hence Lc(X) is regular. Finally, let I be an ideal in Lc(X), generated by {fi}∞i=1, i.e., I = ∑∞ i=1 fiLc(X). If I = C(X), then we are done in this case, for ∞⋂ i=1 Zl(fi) = ∅. Hence we assume that I 6= C(X). Since Lc(X) is regular, we infer that I = ∑∞ i=1 ⊕eiLc(X), where each ei, i ∈ I is an idempotent in Lc(X), and for each i 6= j, eiej = 0, see [10, Theorem 5.5], or [17, Lemma 2], [8, Proposition 1.4]. If x ∈ X, ej(x) 6= 0, then for each i 6= j, ei(x) = 0, and ⋂ Z[I] = ∞⋂ i=1 Z(fi) = ∞⋂ i=1 Z(ei) Now, we may define g = ∑∞ i=1 ei pi(1+ei) , where p ≥ 2 is a real number. Clearly g ∈ C(X), and Z(g) = ⋂∞ i=1 Z(ei). On the other hand for each x ∈ X, there exists at most a unique i ≥ 1 such that ei(x) 6= 0. Therefore g(x) = ei(x) pi(1+ei(x)) = 1 2pi . Hence g(X) ⊆ {0, 1 2p , 1 2p2 , . . .} i.e., g ∈ Cc(X), therefore g ∈ Lc(X). � Remark 4.13. In view of the previous proof we may record an interesting fact, which follows. Let X be a LCP-space and {fi}i∈I be an infinite countable set of elements in C(X), then ⋂ i∈I Z(fi) = Z(g), where g ∈ Cc(X) ⊆ Lc(X) can be chosen with the property that g(X) is an infinite subset of an arbitrary subfield of R. It is well known that X is a P-space if and only if every Gδ-set is open, see [13, 4J(3)]. The following theorem is the counterpart of this result, see also [10, Corollary 5.7]. Corollary 4.14. Let X be a lc-completely regular LCP-space. Then every Gδ-set A containing a compact set S contains a zero-set in Zl(X) containing S. In particular, every lc-completely regular LCP-space is a P-space. If Mlp = Mp ∩Lc(X) and Olp = Op ∩Lc(X), where p ∈ X and Op is the ideal of C(X) consisting of all f in C(X) for which Z(f) is a neighborhood of p. It goes without saying that Mlp is a maximal ideal in Lc(X) and O l p is a zl-ideal in Lc(X). The following theorem is the counterpart of [13, 4J], [10, Theorem 5.8]. c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 196 On the locally functionally countable sub algebra of C(X) Theorem 4.15. Let X be a topological space. Then the following statements are equivalent. (1) X is a LCP-space. (2) Lc(X) is a regular ring. (3) Each ideal in Lc(X) is a zl-ideal. (4) Each prime ideal in Lc(X) is a maximal ideal. (5) For each p ∈ X, Mlp = Olp. (6) Every zero-set in Zl(X) is open. (7) Each ideal in Lc(X) is an intersection of maximal ideals. (8) For all f, g ∈ Lc(X), (f, g) = (f2 + g2). (9) For every f ∈ Lc(X), Zl(f) (X \ Zl(f)) is C-embedded. (10) If {fi : i ∈ N} ⊆ Lc(X), then ⋂∞ i=1 Zl(fi) is an open zero-set in Zl(X). The following results are the counterparts of [11, Proposition 2.5] and [11, Corollary 2.6]. Proposition 4.16. Lc(X) is regular if and only if every pseudoprime ideal in Lc(X) is prime. Corollary 4.17. Let X be a lc-completely regular. Then every pseudoprime ideal in C(X) is prime if and only every pseudoprime ideal in Lc(X) is prime. The following theorem is similar to [13, Theorem 4.11], [11, Theorem 3.8]. Theorem 4.18. Let X be a lc-completely regular space, then the following statements are equivalent. (1) X is compact. (2) Every ideal of Lc(X) is fixed. (3) Every maximal ideal of Lc(X) is fixed. (4) Every prime ideal of Lc(X) is fixed. If X is any topological space and x ∈ X, Mlx = Mx ∩ Lc(X), then as we pointed out earlier Mlx is a maximal ideal of Lc(X) and in fact Lc(X) Mlx ∼= R. Consequently, the Jacobson radical of Lc(X) is zero. Definition 4.19. A maximal ideal M in Lc(X) is called a real maximal ideal of Lc(X) if Lc(X) M ∼= R. A topological space X is called locally countably realcompact space (briefly, lc-realcompact) if every real maximal ideal M of Lc(X) is of the form M = M l x for some x ∈ X. The following results are the counterparts of [13, 10.5(c)] and [11, Theorem 3.11]. Theorem 4.20. X is a lc-realcompact space if and only if each nonzero ho- momorphism from Lc(X) into R is a valuation map. If X is a compact zero-dimensional space, the corresponding x → Mlx is one-one from X onto the set of maximal ideals of Lc(X), say Max(Lc(X)), and hence the space X is homeomorphic to Max(Lc(X)) with the Stone topology c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 197 O. A. S. Karamzadeh, M. Namdari and S. Soltanpour (note, the proof is similar to [13, 4.9(a)], see also the comment above [10, Theorem 3.9] and [19]). The proof of the following result which is similar to its counterpart in [13, Theorem 8.3], is omitted. Proposition 4.21. Two zero-dimensional lc-realcompact spaces X and Y are homeomorphic if and only if Lc(X) ∼= Lc(Y ). We recall that if X, Y are compact zero-dimensional spaces, then C(X) ∼= C(Y ) if and only if Cc(X) ∼= Cc(Y ). In what follows we show that this result also holds if we replace Cc(X) by Lc(X), but the proof is not as evident. Theorem 4.22. Let X and Y be two lc-completely regular compact spaces (e.g., zero-dimensional compact spaces). Then X and Y are homeomorphic if and only if Lc(X) ∼= Lc(Y ). In particular, if X, Y are compact zero- dimensional spaces, then Lc(X) ∼= Lc(Y ) if and only if Cc(X) ∼= Cc(Y ) if and only if CF (X) ∼= CF (Y ) if and only if C(X) ∼= C(Y ). Proof. Clearly if Lc(X) ∼= Lc(Y ), then Max(Lc(X)) and Max(Lc(Y )) are homeomorphic (with the Stone topology), i.e., X, Y are homeomorphic, see the comment preceding Proposition 4.21. Conversely, let ϕ : X → Y be a home- omorphism from X onto Y . If f ∈ Lc(Y ), then we claim that foϕ ∈ Lc(X). To see this, since f ∈ Lc(Y ), we infer that Y = Cf = ⋃ i∈I Vi, where for each i ∈ I, Vi is open in Y and |f(Vi)| ≤ ℵ0. Let us put Ui = ϕ−1(Vi), where i ∈ I. Clearly Ui is open in X and |foϕ(Ui)| = |foϕ(ϕ−1(Vi))| = |f(Vi)| ≤ ℵ0, hence Cfoϕ ⊇ ⋃ i∈I Ui. Since ϕ is open (note, ϕ −1 is continuous), we infer that X = ϕ−1(Y ) = ϕ−1( ⋃ i∈I Vi) ⊆ ϕ−1( ⋃ i∈I Vi) = ⋃ i∈I ϕ−1(Vi) = ⋃ i∈I Ui Therefore Cfoϕ = X, i.e., foϕ ∈ Lc(X). Now we define σ : Lc(Y ) → Lc(X) with σ(f) = foϕ. It is evident that σ is an isomorphism from Lc(Y ) onto Lc(X). The last part is evident. � Remark 4.23. The above result shows that if X, Y are compact zero-dimensional spaces, such that C(X) ∼= C(Y ), then Lc(X) ∼= Lc(Y ). In the comment fol- lowing [11, Corollary 9.5], it is observed that whenever X, Y are two arbi- trary spaces (not necessary compact zero-dimensional) and C(X) ∼= C(Y ), then Cc(X) ∼= Cc(Y ) and CF (X) ∼= CF (Y ) (i.e., Cc(X) and CF (X) are al- gebraic objects). This naturally raises the question that whether Lc(X) is also an algebraic object, too (i.e., if C(X) ∼= C(Y ), then is Lc(X) ∼= Lc(Y ))? Clearly, if X, Y are strongly zero-dimensional spaces with C(X) ∼= C(Y ), then Lc(βX) ∼= Lc(βY ). Let us recall that a commutative ring R is selfinjective (resp., ℵ0-selfinjective), if every homomorphism f : I → R, where I is an ideal (resp., countably gen- erated ideal) in R, can be extended to f̂ : R → R. We recall that a subset S of a commutative ring R is said to be orthogonal, provided xy = 0 for all x, y ∈ S with x 6= y. In the following result we show that [10, Theorem 6.10] is also true for Lc(X). In contrast to the proofs of some of the previous results, c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 198 On the locally functionally countable sub algebra of C(X) we should emphasize that the next proof can not be easily obtained from the proof of its counterpart (i.e., [10, Theorem 6.10]). It is well known that the ℵ0-selfinjectivity of a ring is not a consequence of its regularity, in general, see [14, Examples 14.7, 14.9]. But, the following worthwhile fact shows that Lc(X) as well as C(X) and Cc(X) have this rare property, see [8], [10]. We should remind the reader that CF (X) does not satisfy this property in general, see [10, Remark 6.11, Example 7.1] (note, CF (X) is always regular, see [11, the comment preceding Proposition 4.2]. Theorem 4.24. Let X be a topological space. Then Lc(X) is regular if and only if Lc(X) is ℵ0-selfinjective. Proof. If Lc(X) is ℵ0-selfinjective, then Lc(X) is regular by [18, Proposition 1.2], or [10, Lemmas 6.7, 6.8, Remark 6.9]. Conversely, by [18, Lemma 1.9] and [10, Lemma 6.8, Remark 6.9], it suffices to show that if S is an orthogonal subset in Lc(X), then there exists f ∈ Lc(X) such that for each g ∈ S, fg = g2. Let S = {fi}∞i=1, where fi 6= 0, for each i ∈ I. Since Lc(X) is regular, ⋂∞ i=1 Z(fi) = Z(h) is an open zero-set in Lc(X), by Theorem 4.12. Put Gi = X\Z(fi), for each i ≥ 1. Since fifj = 0, hence Gi ∩ Gj = ∅, for each i 6= j, and Gi’s are clopen for each i ≥ 1. Let us put G = ⋃∞ i=1 Gi, hence X = ⋃∞ i=1 Gi ∪ (X\G). We may define f : X → R by f(x) = { fi(x) , x ∈ Gi 0 , x /∈ G i.e., f|Gi = fi for all i ≥ 1 and f(x) = 0 for all x ∈ X\G. Hence f is continuous by [13, 1A(2)] and we must show that f ∈ Lc(X). Let V ⊆ X be an arbitrary open set, then we are to show that there exists an open set U in X such that |f(U)| ≤ ℵ0 and U ∩ V 6= ∅. Now we consider two cases. First let V ⊆ X\G, then f(V ) = 0, hence V ⊆ Cf . Otherwise V ∩ G 6= ∅, hence there exists a nonempty open subset Gi such that V ∩ Gi 6= ∅. Since fi ∈ Lc(X) i.e., Cfi = X, hence there exists an open set H ⊆ Cfi such that |fi(H)| ≤ ℵ0 and ∅ 6= H ∩ (V ∩ Gi) = U. Now clearly, |f(U)| = |fi(U)| ≤ |fi(H)| ≤ ℵ0 i.e., we are done. Finally, we claim that ffi = f 2 i , for each fi ∈ S and this complete the proof, by [18, Lemma 1.9]. To this end, we note that if f(x) = 0, then x /∈ G, hence x /∈ Gi for all i ≥ 1, i.e., x ∈ Z(fi), for all i ≥ 1. Thus ffi = f2i , on Z(f) for each fi ∈ S. Since f(x) = fi(x), for each x ∈ Gi = X\Z(fi) and Z(f) ⊆ Z(fi) for each i ≥ 1, we infer that ffi = f2i , for each fi ∈ S, hence we are done. � Remark 4.25. Let X be an uncountable discrete space, then C(X) = Lc(X) is selfinjective but Cc(X) is not selfinjective, see [10, Example 7.1, Remark 7.5]. More generally, if C(X) is ℵ0-selfinjective, then by [8, Theorem 1], X is a P-space. Hence in view of Proposition 2.18, we have L1(X) = LF (X) = Lc(X) = C(X). Moreover in view of Theorem 4.22 and Remark 4.11, we note that the ℵ0-selfinjectivity of C(X), Lc(X), Cc(X), and C(X, K) coincide if X is a zero-dimensional space. c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 199 O. A. S. Karamzadeh, M. Namdari and S. Soltanpour 5. The socle of Lc(X) We recall that the socle of any commutative ring R, Soc(R), is the sum of its nonzero minimal ideals (in fact, it can be written as the direct sum of some of its nonzero minimal ideals). We recall that CF (X) = {f ∈ C(X) : |X\Z(f)| < ∞}, the socle of C(X), is a z-ideal. Clearly, CF (X) ⊆ LF (X) ⊆ Lc(X). We show that CF (X) ⊆ Soc ( Lc(X) ) , i.e., CF (X) is a sum of minimal ideals of Lc(X). In [20, Proposition 3.1], it is shown that if I is a minimal ideal of C(X), then I = eC(X), where e is an idempotent such that e(x) = 1 and e(X\{x}) = 0, where x is an isolated point of X. Clearly C(X) = eC(X) ⊕ (1 − e)C(X), eC(X) = eLc(X) = eCc(X), and Lc(X) = eLc(X) ⊕ (1 − e)Lc(X). We also note that (1 − e)Lc(X) = (1 − e)C(X) ∩ Lc(X) is a maximal ideal in Lc(X), therefore eLc(X) = eC(X) is a minimal ideal in Lc(X). Hence every minimal ideal of C(X) is a minimal ideal in Lc(X), too. Therefore CF (X) is an ideal in Lc(X), and CF (X) ⊆ Soc ( Lc(X) ) . We should also emphasize that since Soc ( Lc(X) ) is a semisimple Lc(X)-module, hence CF (X) is a direct summand of Soc ( Lc(X) ) as a Lc(X)-module. In the proof of Theorem 5.4, we shall briefly observe that Soc ( Lc(X) ) ⊆ Soc ( Cc(X) ) . Let us also recall that CF (X) is the subring of C(X) whose elements have finite image. Hence, we have CF (X) ⊆ CF (X) ⊆ Cc(X) ⊆ Lc(X) ⊆ C(X) and Soc ( Cc(X) ) = Soc ( CF (X) ) , see [10], [11]. The following lemma which is similar to [11, Lemma 5.1], some- how determines the minimal ideals of Lc(X). Let us first remind the reader that if I is a nonzero minimal ideal in a reduced commutative ring R, then I = eR, where e ∈ R is an idempotent (note, I = (a) = (a2), for every 0 6= a ∈ I, and a = a2r, for some r ∈ R, now put e = ar). In [11, Lemma 5.1], it is shown that if 0 6= e is an idempotent, then eCc(X) is a minimal ideal in Cc(X) if and only if Z(1 − e) is connected. In the next lemma the minimal ideals in Lc(X) are characterized, too. Lemma 5.1. Let I be a nonzero minimal ideal in Lc(X), then I = eLc(X) where e is an idempotent in Lc(X) such that Z(1−e) is connected. Conversely, if I = eLc(X) where e 6= 0 is an idempotent in Lc(X) such that Z(1 − e) is a constant subset of X with respect to Lc(X), then I is a minimal ideal in Lc(X). Proof. Let I be a nonzero minimal ideal in Lc(X). Since Lc(X) is reduced, I = eLc(X), where e is an idempotent in Lc(X). If Z(1 − e) is not connected, there exists a nonempty clopen subset A ( Z(1 − e) (note, A is clopen in X, too). Now define the idempotent e1 ∈ Lc(X) such that A = Z(1 − e1). Clearly Z(1 − e1) ( Z(1 − e). Consequently, e1 = ee1 but e 6= e1e. Hence e1Lc(X) ( eLc(X) = I and this contradicts the minimality of I. Conversely, let I = eLc(X), where e ∈ Lc(X) such that Y = Z(1 − e) ⊆ X is a constant subset of X with respect to Lc(X) (i.e., Lc(X) ⊆ C1(Y ), see Definition 3.16). We are to show that I is minimal in Lc(X). It suffices to show that (1−e)Lc(X) is a maximal ideal in Lc(X). Now we define ϕ : Lc(X) → R by ϕ(f) = f(Y ). c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 200 On the locally functionally countable sub algebra of C(X) Clearly, kerϕ = (1 − e)Lc(X) and Lc(X)(1−e)Lc(X) ∼= R, hence (1 − e)Lc(X) is maximal in Lc(X). � In [11, Proposition 5.3], the socle of Cc(X) is characterized and in [11, Remark 5.2, and the introduction of Section 5], it’s observed that CF (X) ⊆ Soc ( Cc(X) ) = Soc ( CF (X) ) . Next, we topologically characterize the socle of Lc(X). The next proof is similar to the proof of [11, Proposition 5.3], but it’s given for the sake of the reader. Proposition 5.2. Let f ∈ Lc(X) be a nonunit element. If f ∈ Soc ( Lc(X) ) , then X\Z(f) ⊆ ⋃n i=1 Ai, where n ∈ N and {A1, A2, . . . , An} is a set of mutually disjoint clopen connected subsets of X. Conversely, if X\Z(f) ⊆ ⋃n i=1 Ai, where n ∈ N and {A1, A2, . . . An} is a set of mutually disjoint clopen constant subsets of X with respect to Lc(X), then f ∈ Soc ( Lc(X) ) . In Particular, Soc ( Lc(X) ) is a zl-ideal in Lc(X). Proof. We put Soc ( Lc(X) ) = ∑ i∈I ⊕eiLc(X), where each ei is an idempotent in Lc(X), and eiLc(X) is a nonzero minimal ideal in Lc(X). Let f = ei1f1 + ei2f2 + . . .+ einfn be an element in Soc ( Lc(X) ) , where fk ∈ Lc(X) and ik ∈ I, k = 1, 2, . . . , n. We put Aik = Z(1 − eik), for each ik ∈ I, k = 1, 2, . . . , n. Clearly, Aik , k = 1, 2, . . . , n, are clopen and connected, by Lemma 5.1. Since the idempotent elements {ei : i ∈ I} are mutually orthogonal, we infer that {Ai : i ∈ I, Ai = Z(1 − ei)} is a set of mutually disjoint clopen connected subsets of X. If x /∈ ⋃n i=1 Ai, then eik (x) = 0, k = 1, 2, . . . , n, hence x ∈ Z(f). Therefore X\Z(f) ⊆ ⋃n i=1 Ai. Conversely, let X\Z(f) ⊆ ⋃n i=1 Ai, where {Ai : i ∈ I} is a set of mutually disjoint clopen constant subsets in X with respect to Lc(X), we show that f ∈ Soc ( Lc(X) ) . Since each Ai is a clopen set, there exists an idempotent ei, such that Ai = Z(1 − ei), where i = 1, 2, . . . , n. We also note that each Ai is constant with respect to Lc(X), hence there is a set of idempotents in Lc(X), {e1, . . . , en} say, which are mutually orthogonal and each eiLc(X) is a minimal ideal in Lc(X), by Lemma 5.1. Clearly, f = e1f + e2f + . . . + enf ∈ Lc(X) which belongs to Soc ( Lc(X) ) = ∑ i∈I ⊕eiLc(X). � Remark 5.3. One can easily observe that if in the previous two results we trade off Lc(X) with any R-subalgebra of Lc(X), A say, which contains Cc(X), then the two results are also valid for A. The next result determines spaces X such that the socles of Lc(X), Cc(X) and hence of CF (X) coincide. Theorem 5.4. Soc ( Lc(X) ) = Soc ( Cc(X) ) if and only if the clopen connected subsets of X coincide with the clopen constant subsets of X with respect to Lc(X). Proof. Soc ( Lc(X) ) ⊆ Soc ( Cc(X) ) , for if I is a minimal ideal in Soc ( Lc(X) ) , then I = eLc(X) where e is an idempotent such that Z(1 − e) is connected, by Lemma 5.1. Hence I is a minimal ideal in Cc(X), by [11, Lemma 5.1]. Now, let I be a nonzero minimal ideal in Soc ( Cc(X) ) , so I = eCc(X), where e 6= 0, 1 is c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 201 O. A. S. Karamzadeh, M. Namdari and S. Soltanpour an idempotent and Z(1 − e) is a clopen connected subset in X, by [11, Lemma 5.1]. Hence by our hypothesis Z(1 − e) is constant with respect to Lc(X). Therefore I = eLc(X) = eCc(X) is a minimal ideal of Lc(X), by Lemma 5.1, hence it is in Soc ( Lc(X) ) . Conversely, let Soc ( Lc(X) ) = Soc ( Cc(X) ) , and ∅ 6= Y ⊆ X be a clopen constant subspace of X with respect to Lc(X) we are to show that Y is connected. Clearly, there exists e ∈ Lc(X) such that e(Y ) = 1, e(X\Y ) = 0. But Y = Z(1 − e), hence by Lemma 5.1, we infer that e ∈ Soc ( Lc(X) ) = Soc ( Cc(X) ) . Consequently, Y = Z(1 − e) must be connected, by [11, Proposition 5.3]. � We remind the reader in [11, Theorem 6.6], it is proved that Soc ( Cc(X) ) = CF (X) if and only if each clopen connected subsets of X consists of a single isolated point. Motivated by this fact and Theorem 5.4, we present the next result. Theorem 5.5. If every proper nonempty clopen connected subset of X is sin- gleton, (e.g., any totally disconnected space), then Soc ( Lc(X) ) = CF (X). Con- versely, if Soc ( Lc(X) ) = CF (X), then every proper nonempty clopen constant subspace of X with respect to Lc(X) is singleton. Proof. Let every proper nonempty clopen connected subset of X be single- ton, we are to show that Soc ( Lc(X) ) = CF (X). It is evident that CF (X) ⊆ Soc ( Lc(X) ) . Let I be a nonzero minimal ideal in Soc ( Lc(X) ) , so by Lemma 5.1, I = eLc(X), where e 6= 0, 1 is an idempotent and Z(1 − e) is a clopen connected subset in X. Hence by our hypothesis Z(1 − e) is singleton. There- fore I = eLc(X) = eC(X) is a minimal ideal in CF (X), by [20, Proposition 3.3]. Conversely, let CF (X) = Soc ( Lc(X) ) , and ∅ 6= Y ⊆ X be a clopen con- stant subspace of X with respect to Lc(X). There exists e ∈ Lc(X) such that e(Y ) = 1, e(X\Y ) = 0. Clearly, by Lemma 5.1, e ∈ Soc ( Lc(X) ) = CF (X), hence eC(X) is a minimal ideal in CF (X), therefore Y = Z(1 − e) is single- ton. � The following remark is now immediate. Remark 5.6. CF (X) = Soc ( Lc(X) ) = Soc ( Cc(X) ) if and only if each clopen connected subset of X consists of a single isolated point. Consequently, if X is zero-dimensional or totally disconnected, we have CF (X) = Soc ( Lc(X) ) = Soc ( Cc(X) ) . Let us recall that an ideal in a commutative ring R is essential if it intersects every nonzero ideal of R nontrivially. It is well known and easy to show that a nonzero ideal I in a reduced ring R (i.e., no nonzero element in R is nilpotent) is essential if and only if Ann(I) = 0, see [3, Background and preliminary results]. The proof of the following corollary is similar to [11, Corollary 5.4], but we include the proof for the sake of the reader. Corollary 5.7. Let X be a lc-completely regular space, and Soc ( Lc(X) ) =∑ i∈I ⊕eiLc(X), where eiLc(X) is a nonzero minimal ideal of Lc(X), and ei c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 202 On the locally functionally countable sub algebra of C(X) is an idempotent for each i ∈ I. Put Y = ⋃ i∈I Z(1 − ei), then Soc ( Lc(X) ) is essential in Lc(X) if and only if Y is dense in X. Proof. Let Y = ⋃ i∈I Z(1−ei) be dense in X, we are to show that Soc ( Lc(X) ) is essential in Lc(X). Since Lc(X) is reduced, in order to prove that Soc ( Lc(X) ) is essential in Lc(X) it suffices to show that Ann(Soc ( Lc(X) ) ) = (0). We note that f ∈ Ann(Soc ( Lc(X) ) ) if and only if fei = 0 for each i ∈ I. Now, if fei = 0, then f(Z(1 − ei)) = 0, hence f(Y ) = {0}. Since Y is dense in X we infer that f = 0, and we are done. Conversely, let Soc ( Lc(X) ) be essential in Lc(X), hence Ann(Soc ( Lc(X) ) ) = (0) in Lc(X). Let us now take x ∈ X\Y and obtain a contradiction. By lc-complete regularity of X, there exists 0 6= f ∈ Lc(X) with f(Y ) = f(Y ) = 0. Therefore f(Z(1 − ei)) = 0, hence fei = 0 for all i ∈ I. Thus 0 6= f ∈ Ann(Soc ( Lc(X) ) ) = (0), which is a contradiction. � We recall that CF (X) is never a prime ideal of C(X), see [8, Proposition 1.2], or [3, Remark 2.4]. The following result characterizes spaces X such that CF (X) 6= 0 is a prime ideal in Lc(X) (note, CF (X) 6= 0 if and only if X has isolated points). Proposition 5.8. Let |I(X)| < ∞, where I(X) is the set of isolated points in X. If 0 6= CF (X) is a prime ideal in Lc(X), then X\I(X) is connected in X. Conversely, if X\I(X) is constant with respect to Lc(X), then 0 6= CF (X) is prime in Lc(X). Proof. Let Y = X\I(X) = A ∪ B, where A, B are two nonempty infinite disjoint clopen subsets of Y and seek a contradiction. Since Y is clopen in X we infer that A, B are also clopen in X. Clearly, X = I(X) ∪ A ∪ B. Now define f, g ∈ Lc(X) such that f(A∪I(X)) = 1, f(B) = 0 and g(A∪I(X)) = 0, g(B) = 1. Clearly fg = 0 ∈ CF (X), but by [20, Proposition 3.3], we infer that f, g /∈ CF (X), which is a contradiction. Conversely, let Y = X\I(X) be constant with respect to Lc(X) and take f, g ∈ Lc(X) such that fg ∈ CF (X). Clearly X = Y ∪ I(X), so X\Z(fg) ⊆ I(X) and fg(Y ) = 0. Since f and g are constant on Y , we infer that either f(Y ) = 0 or g(Y ) = 0, i.e., X\Z(f) ⊆ I(X) or X\Z(g) ⊆ I(X), therefore f ∈ CF (X) or g ∈ CF (X), by [20, Proposition 3.3], and we are done. � In the following corollary, we consider spaces X, such that CF (X) is not a prime ideal in Lc(X). Corollary 5.9. If I(X) is an infinite set or Y = X\I(X) is disconnected, then CF (X) is never a prime ideal in Lc(X). Proof. Let I(X) be an infinite set and take A = {xn : n ∈ N}, B = {yn : n ∈ N} to be two disjoint countably infinite subsets of I(X). We now define f(x) = { 1 n , x = xn ∈ A 0 , x /∈ A and g(x) = { 1 n , x = yn ∈ B 0 , x /∈ B . Let ǫ > 0 be given, then there exists k ∈ N such that 1 n < ǫ, for all n ≥ k. Now, for the clopen subsets G = X\{x1, x2, . . . , xk}, H = X\{y1, y2, . . . , yk} and c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 203 O. A. S. Karamzadeh, M. Namdari and S. Soltanpour for each x ∈ G, y ∈ H, we have |f(x)| < ǫ, |g(y)| < ǫ, hence f, g ∈ C(X). Clearly, f, g ∈ Cc(X). Therefore f, g ∈ Lc(X) and 0 = fg ∈ CF (X), but f, g /∈ CF (X), by [20, Proposition 3.3]. Consequently, in this case CF (X) is not prime in Cc(X), a fortiori, in Lc(X)). Finally let |I(X)| < ∞ and X\I(X) be disconnected, hence by Proposition 5.8, we are done. � In the next result, which is our main theorem in this section, we consider the maximality of CF (X) in Lc(X). First, let us recall that if ϕ : C(X) → C(Y ) is a ring homomorphism with ϕ(1) = 1, then ϕ(Cc(X)) ⊆ Cc(Y ). This is an easy consequence of the fact that whenever f ∈ Cc(X), then Im(ϕ(f)) ⊆ Im(f) (note, let r ∈ Im(ϕ(f)), then ϕ(f) − r is non-unit, but ϕ(f − r) = ϕ(f) − r, hence f − r is non-unit too, i.e., Z(f − r) 6= ∅, and we are done), see also[11, the comment following Corollary 3.5]. Theorem 5.10. Let CF (X) be a maximal ideal in Lc(X). Then C F (X) = Cc(X) = Lc(X) and Lc(X) is isomorphic to a finite direct product of fields, each of which, is isomorphic to R and X has a unique infinite clopen connected subset. Conversely, let X have a unique infinite clopen connected subset, and assume that every element of Lc(X) is constant on it, and Lc(X) ∼= n∏ i=1 Fi, where each Fi is a field. Then C F (X) = Cc(X) = Lc(X), and CF (X) is maximal in Lc(X). Proof. Let CF (X) be a maximal ideal in Lc(X). Let us first take care of the case, when CF (X) = 0. Clearly in this case Lc(X) = R (note, in this case X is connected and Cc(X) = C F (X) = R), and we are done. Hence, we may assume that that CF (X) 6= 0. In view of the previous corollary we infer that I(X), the set of isolated points of X must be finite. Let us assume that |I(X)| = n, where n is a positive integer (note, CF (X) 6= 0 if and only if I(X) 6= ∅, see [20, Proposition 3.3]). Hence CF (X) is a finitely generated ideal in C(X) (note, by [20, Proposition 3.1], there is a one-one correspondence between I(X) and the set of nonzero minimal ideals in C(X)). Consequently, CF (X) = n∑ i=1 ⊕eiC(X), where each ei is an idempotent and eiC(X) = eiLc(X) is a minimal ideal in C(X) as well as in Lc(X), see the comment preceding Lemma 5.1. Clearly, CF (X) = eCF (X) = eC(X) = eLc(X), where e = e1 + e2 + · · · + en (note, eiej = 0 for i 6= j). Since CF (X) is maximal in Lc(X), we infer that e 6= 1, which implies that (1−e)Lc(X) is a nonzero minimal ideal in Lc(X). Inasmuch as CF (X) is maximal in Lc(X) and CF (X) ⊆ Soc ( Lc(X) ) , we infer that either CF (X) = Soc ( Lc(X) ) or Lc(X) = Soc ( Lc(X) ) . We claim that CF (X) = Soc ( Lc(X) ) leads us to a contradiction. To see this, we note that (1−e)Lc(X) is a nonzero minimal ideal in Lc(X). Hence if the latter equality holds, we infer that (1 − e)Lc(X) must be in CF (X). But CF (X) ∩ (1 − e)Lc(X) = 0, which is absurd. Consequently, we must have Lc(X) = Soc ( Lc(X) ) = CF (X) ⊕ (1 − e)Lc(X). Now, for each ei we can easily show that eiLc(X) ∼= R ∼= (1−e)Lc(X). To see this, let x ∈ Z(1 − ei) and define ϕ : Lc(X) → R by ϕ(f) = f(x) for c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 204 On the locally functionally countable sub algebra of C(X) all f ∈ Lc(X). Hence, (1 − ei)Lc(X) ⊆ kerϕ. Since (1 − ei)Lc(X) is maximal in Lc(X), we infer that (1 − ei)Lc(X) = kerϕ, hence eiLc(X) ∼= Lc(X)(1−ei)Lc(X) ∼= R. Similarly (1 − e)Lc(X) ∼= Lc(X)eLc(X) ∼= R. Consequently, we have already shown that Lc(X) ∼= n+1∏ i=1 Ri, where Ri = R. In view of Lemma 5.1 , and by the fact that (1 − e)Lc(X) is minimal in Lc(X), we infer that Z(e) is connected. Consequently, by the comment preceding Lemma 5.1, (1 − e)Cc(X) is a minimal ideal in Cc(X). Hence Cc(X) = CF (X) ⊕ (1 − e)Cc(X), which is equal to Soc ( Cc(X) ) = Soc ( CF (X) ) ⊆ CF (X), see the comment preceding Proposition 5.2. Thus CF (X) = Cc(X). But, Cc(X) = CF (X) ⊕ (1 − e)Cc(X) is the direct sum of n + 1 minimal ideals in Cc(X), hence by the above proof for Lc(X), we can also show that Cc(X) ∼= n+1∏ i=1 Ri, where Ri = R. Let us consider the natural isomorphism ϕ : Cc(X) → Lc(X) ⊆ C(X). Now in view of the comment preceding the theorem we have Lc(X) ⊆ Cc(X), hence Lc(X) = Cc(X) = C F (X). Finally, in view of [20, Proposition 3.3], it is clear that the connected clopen set Z(e) is infinite (in fact Z(e) = X \ I(X)). It is also manifest that every non-singleton connected subset of X must be a subset of Z(e), hence Z(e) is the only clopen connected subset of X which is infinite, and we are done. Conversely, since Lc(X) ∼= n∏ i=1 Fi, where each Fi is a field, we infer that Lc(X) = n∑ i=1 ⊕uiLc(X) = Soc ( Lc(X) ) , where each uiLc(X) is a nonzero minimal ideal in Lc(X), and each ui is idempotent with 1 = u1 +u2 · · · un. Now let 1 6= u ∈ Lc(X) be an idempotent such that Z(1−u) is the unique infinite clopen subset of X, on which, every element of Lc(X) is constant. Consequently, uLc(X) is a minimal ideal in Lc(X), by Lemma 5.1. Multiplying, 1 = u1 + u2 · · · un by u, we get u = uu1 + uu2 + · · · uun. Clearly, u 6= 0, hence uui 6= 0 for some i. We now claim that there is a unique i, with 1 ≤ i ≤ n such that uui 6= 0. To see this, let uui 6= 0 6= uuj for some i 6= j and obtain a contradiction. But uui 6= 0 implies that uLc(X)uiLc(X) 6= 0, hence uLc(X)uiLc(X) = uLc(X) = uiLc(X) and similarly uLc(X) = ujLc(X), which is a contradiction. Consequently, we may assume that uui = 0 for 1 ≤ i ≤ n − 1 and uun 6= 0. This means that uLc(X) = unLc(X). In view of [20, Proposition 3.3], and the fact that Z(1 − u) is infinite, we infer that u /∈ CF (X), i.e, un /∈ CF (X). By Lemma 5.1, and the fact that each uiLc(X) for 1 ≤ i ≤ n − 1 is minimal, we infer that each Z(1 − ui) is connected, which by our assumption is not an infinite set, hence it must be a singleton. Consequently, in view of [20, Proposition 3.3], ui ∈ CF (X) for 1 ≤ i ≤ n − 1. Inasmuch as Lc(X) = n∑ i=1 ⊕uiLc(X) = Soc ( Lc(X) ) , we infer that CF (X) = n−1∑ i=1 ⊕uiLc(X) ⊕ unLc(X) ⋂ CF (X). Since unLc(X) is minimal, we infer that either unLc(X) ⋂ CF (X) = 0 or unLc(X) ⋂ CF (X) = unLc(X). The latter c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 205 O. A. S. Karamzadeh, M. Namdari and S. Soltanpour equality is impossible, for by what we have already observed above un /∈ CF (X). Consequently, CF (X) = n−1∑ i=1 ⊕uiLc(X) and Lc(X) = CF (X)⊕unLc(X), hence CF (X) is maximal in Lc(X). Now by the proof of the first part we also have CF (X) = Cc(X) = Lc(X), hence we are done. � The following theorem shows that for the spaces X in which there exist certain constant subsets with respect to Lc(X), Lc(X) can not be isomorphic to any C(Y ). Theorem 5.11. Let |I(X)| < ∞ and X\I(X) be constant with respect to Lc(X) (note, in this case Lc(X) = C F (X)). Then there is no space Y with Lc(X) ∼= C(Y ). Proof. Let |I(X)| < ∞ and X\I(X) be constant with respect to Lc(X), then CF (X) is a prime ideal in Lc(X) by Proposition 5.8. If there exists a space Y such that Lc(X) ∼= C(Y ), then Soc ( Lc(X) ) ∼= CF (Y ). Now, since Soc ( Lc(X) ) is a zl-ideal containing a prime ideal CF (X), Soc ( Lc(X) ) is a prime ideal in Lc(X), by Theorem 3.14. Hence CF (Y ) is a prime ideal in C(Y ), which is a contradiction, see the comment preceding Proposition 5.11. � Remark 5.12. If we replace Lc(X) by LF (X) or by L1(X) in this section, then some of the results of this section remain valid for these two rings, too. Remark 5.13. Let X = W ∪{x1, x2, . . . , xn}, where W is constant with respect to L1(X) (e.g., if we take W as in Remark 2.10) and x1, x2, . . . , xn are the only isolated points of X (note, W is connected and has no isolated point) i.e., |I(X)| < ∞ and X\I(X) = W is a constant subset of X with respect to L1(X). Hence, by Theorem 5.11, Remark 5.12, L1(X) can not be isomorphic to any C(Y ), in general. But in some special cases, namely, L1(W) and L1(X) we have L1(W) = R and L1(X) ∼= ∏n i=1 Ri, where Ri = R, for i = 1, 2, . . . , n. That is to say L1(W) = C(Y ), where Y is a singleton, and L1(X) = C(Z), where |Z| < ∞. But, we should remind the reader that we are interested in infinite spaces. Remark 5.14. Let K be a subring of R, then Lc(X, K) is a subring of Lc(X) whose elements take values in K. We denote Lc(X, Z), Lc(X, Q) by Li(X) and Lr(X), respectively. Clearly, Li(X) = C(X, Z) = Ci(X) and Lr(X) = C(X, Q) = Cr(X), see also [10, the comment following Definition 2.1]. It is manifest that Li(X) ⊆ Lr(X) ⊆ C(X, F) ⊆ Cc(X) ⊆ Lc(X), where F is a countable subfield of R and Li(X) ⊆ Lr(X) ⊆ Lc(X, K) ⊆ Lc(X), where K is a proper subfield of R. But unfortunately, apart from Cc(X) and Lc(X), these are not R-subalgebras of C(X), see [10, Remark 7.5], and are not of our interest, in general. c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 206 On the locally functionally countable sub algebra of C(X) References [1] F. Azarpanah, Intersection of essential ideals in C(X), Proc. Amer. Math. Soc. 125 (1997), 2149–2154. [2] F. Azarpanah and O. A. S. Karamzadeh, Algebraic characterization of some discon- nected spaces, Italian. J. Pure Appl. Math. 12 (2002), 155–168. [3] F. Azarpanah, O. A. S. Karamzadeh and S. Rahmati, C(X) VS. C(X) modulo its socle, Colloq. Math. 3 (2008), 315–336. [4] P. Bhattacharjee, M. L. Knox and W. Wm. Mcgovern, The classical ring of quotients of Cc(X), Appl. Gen. Topol. 15, no. 2 (2014), 147–154. [5] O. Dovgoshey, O.Martio, V. Ryazanov and M. Vuorinen, The Cantor function, Expo. Math. 24 (2006), 1–37. [6] T. Dube, Contracting the Socle in Rings of Continuous Functions, Rend. Semin. Mat. Univ. Padova. 123 (2010), 37–53. [7] R. Engelking, General Topology, Heldermann Verlag Berlin, 1989. [8] A. A. Estaji and O. A. S. Karamzadeh, On C(X) modulo its socle, Comm. Algebra 31 (2003), 1561–1571. [9] B. R. Gelbaum and J. M. H. Olmsted, Counterexamples in Analysis, Dover Holden-Day, Inc., San Francisco, (2003). [10] M. Ghadermazi, O. A. S. Karamzadeh and M. Namdari, On the functionally countable subalgebra of C(X), Rend. Sem. Mat. Univ. Padova, 129 (2013), 47–69. [11] M. Ghadermazi, O. A. S. Karamzadeh and M. Namdari, C(X) versus its functionally countable subalgebra, submitted. [12] S. G. Ghasemzadeh, O. A. S. Karamzadeh and M. Namdari, The super socle of the ring of continuous functions, Mathematica Slovaka, to appear. [13] L. Gillman and M. Jerison, Rings of continuous functions, Springer-Verlag, 1976. [14] K. R. Goodearl, Von Neumann Regular Rings, Pitman, 1979. [15] J. Hart and K. Kunen, Locally constant functions, Fund. Math. 150 (1996), 67–96. [16] M. Henriksen, R. Raphael and R. G. Woods, SP -scattered spaces; a new generalization of scattered spaces, Comment. Math. univ. carolin 48, no. 3 (2007), 487–505. [17] O. A. S. Karamzadeh, On a question of Matlis, Comm. Algebra 25 (1997), 2717–2726. [18] O. A. S. Karamzadeh and A. A. Koochakpour, On ℵ 0 -selfinjectivity of strongly regular rings, Comm. Algebra 27 (1999), 1501–1513. [19] O. A. S. Karamzadeh, M. Namdari and M. A. Siavoshi, A note on λ-compact spaces, Math. Slovaca. 63, no. 6 (2013) 1371–1380. [20] O. A. S. Karamzadeh and M. Rostami, On the intrinsic topology and some related ideals of C(X), Proc. Amer. Math. Soc. 93 (1985), 179–184. [21] R. Levy and M. D. Rice, Normal P -spaces and the Gδ-topology, Colloq. Math. 47 (1981), 227–240. [22] M. A. Mulero, Algebraic properties of rings of continuous functions, Fund. Math. 149 (1996), 55–66. [23] M. Namdari and A. Veisi, The subalgebra of Cc(X) consisting of elements with countable image versus C(X) with respect to their rings of quotients, Far East J. Math. Sci. (FJMS), 59 (2011), 201–212. [24] M. Namdari and A. Veisi, Rings of quotients of the subalgebra of C(X) consisting of functions with countable image, Inter. Math. Forum, 7 (2012), 561–571. [25] A. Pelczynski and Z. Semadeni, Spaces of continuous functions (III), Studia Mathemat- ica 18 (1959), 211–222. [26] M. E. Rudin and W. Rudin, Continuous functions that are locally constant on dense sets, J. Funct. Anal. 133 (1995), 120–137. [27] W. Rudin, Continuous functions on compact spaces without perfect subsets, Proc. Amer. Math. Soc. 8 (1957), 39–42. [28] D. Rudd, On two sum theorems for ideals of C(X), Michigan Math. J. 17 (1970), 139– 141. c© AGT, UPV, 2015 Appl. Gen. Topol. 16, no. 2 207