Keywords

2010 Mathematics Subject Classification

1 Introduction

Let D be an integral domain. Consider the following two assertions:

(i) If I is an ideal of D whose localizations at the maximal ideals are finitely generated, then I is finitely generated.

(ii) Every \(x\in D-\{0\}\) belongs to only finitely many maximal ideals of D.

While \((ii)\Rightarrow (i)\) is well-known and easy to prove, Bazzoni [3, p. 630] conjectured that the converse is true for Prüfer domains. Recall that D is a Prüfer domain if every finitely generated ideal I of D is locally principal.

Holland et al. [10, Theorem 10] proved Bazzoni’s conjecture for Prüfer domains using techniques from lattice-ordered groups theory and McGovern [14, Theorem 11] proved the same result using a direct ring theoretic approach. Halter-Koch [9, Theorem 6.11] proved Bazzoni’s conjecture for r-Prüfer monoids (see Section 4). Zafrullah [16, Proposition 5] proved Bazzoni’s conjecture for Prüfer v-multiplication domains. Finocchiaro and Tartarone [6, Theorem 4.5] proved Bazzoni’s conjecture for almost Prüfer ring extensions (see Section 3). Recently, Chang and Hamdi [4, Theorem 2.4] proved Bazzoni’s conjecture for almost Prüfer v-multiplication domains (see Section 4).

The purpose of this paper is to prove a Bazzoni-type theorem for multiplicative lattices (see Section 2), thus unifying the results mentioned above (see Sections 3 and 4). Our standard references are [1, 7, 8].

2 Main Result

We use an abstract ideal theory approach, so we work with multiplicative lattices.

Definition 1

A multiplicative lattice is a complete lattice \((L,\le )\) (with bottom element 0 and top element 1) which is also a multiplicative commutative monoid with identity 1 (the top element) and satisfies \(a( \bigvee b_\alpha ) = \bigvee ab_\alpha \) for each \(a,b_\alpha \in L\).

Let L be a multiplicative lattice. The elements in \(L-\{1\}\) are said to be proper. Denote by Max(L) the set of maximal elements of L. For \(x,y\in L\), set \((y : x) =\bigvee \{a \in L ;\ ax \le y\}\).

We recall some standard terminology.

Definition 2

Let L be a multiplicative lattice and let \(x,p \in L\).

(1) p is prime if \(p\ne 1\) and for all \(a,b \in L\), \(ab \le p\) implies \(a \le p\) or \(b \le p\). It follows easily that every maximal element is prime.

(2) x is compact if whenever \(x\le \bigvee _{y\in S} y\) with \(S\subseteq L\), we have \(x\le \bigvee _{y\in T} y\) for some finite subset T of S.

(3) L is a C -lattice if the set \(L^*\) of compact elements of L is closed under multiplication, \(1\in L^*\) and every element in L is a join of compact elements.

(4) x is meet-principal if \(y \wedge zx = ((y : x) \wedge z) x\) for all \(y, z \in L\) (in particular \((y:x)x=x\wedge y\)).

(5) x is join-principal if \(y \vee (z : x) = ((yx \vee z) : x)\) for all \(y, z \in L\) (in particular \((xy:x)=y\vee (0:x)\)).

(6) x is cancellative if for all \(y,z\in L\), \(xy = xz\) implies \(y = z\).

(7) x is CMP (ad hoc name) if x is cancellative and meet-principal.

(8) L is a lattice domain if \((0:a)=0\) for all \(a\in L-\{0\}\).

In the sequel, we work with C-lattices and their localization theory. Let L be a C-lattice. For \(p \in L\) a prime element and \(x\in L\), we set

$$x_p = \bigvee \{a \in L^* ;\ ab \le x \text{ for } \text{ some } b \in L^* \text{ with } b\not \le p\}.$$

Then \(L_p:=\{ x_p;\ x\in L\}\) is again a lattice with multiplication \((x_p,y_p)\mapsto (xy)_p=(x_py_p)_p\), join \(\{ (b_\alpha )_p \}\mapsto (\bigvee (b_\alpha )_p)_p = (\bigvee b_\alpha )_p\) and meet \(\{ (b_\alpha )_p \}\mapsto (\bigwedge (b_\alpha )_p)_p\). The next lemma collects several basic properties.

Lemma 3

Let L be a C-lattice, let \(x,y\in L\) and let \(p \in L\) be a prime element.

(1) \(x_p = 1\) if and only if \(x\not \le p\).

(2) \((x \wedge y)_p = x_p \wedge y_p \).

(3) If x is compact, then \((y : x)_p = (y_p : x_p)\).

(4) \(x=y\) if and only if \(x_m=y_m\) for each \(m\in Max(L)\).

(5) A cancellative element x is CMP if and only if \((y:x)x=x\wedge y\) for all \(y\in L\).

(6) If x is compact, then \(x_p\) is compact in \(L_p\). Conversely, if \(x_p\) is compact in \(L_p\), then \(x_p=c_p\) for some compact element \(c\le x\).

(7) If x and y are CMP elements, then so is xy.

(8) If x is compact, y is CMP and \(x\le y\), then (x : y) is compact.

Proof

For (1–4) see [11, pp. 201–202], for (5) see [15, Lemma 2.10], while (6–7) follow easily from definitions. We prove (8). Note that \((x:y)y=x\wedge y = x\). Suppose that \((x:y)\le \bigvee _{i\in A} z_{i}\). Then \(x=(x:y)y\le \bigvee _{i\in A} z_{i}y\), so \((x:y)y\le \bigvee _{i\in B} z_{i}y\) for some finite subset B of A. Cancel y to get \((x:y)\le \bigvee _{i\in B} z_{i}\). \(\square \)

Say that x and \(y \in L\) are comaximal if \(x\vee y=1\). Clearly, \(x\vee yz=1\) if and only if \(x\vee y=1\) and \(x\vee z=1\). When \(t\le u\) we say that t is below u or that u is above t.

Lemma 4

Let L be a C-lattice and \(z\in L-\{1\}\) a compact element such that \(\{ m\in Max(L);\ z\le m\}\) is infinite. There exists an infinite set \(\{ a_n;\ n\ge 1\}\) of pairwise comaximal proper compact elements such that \(z\le a_n\) for each n.

Proof

We may clearly assume that \(z=0\) (just change L by \(\{ x\in L;\ x\ge z\}\)). Say that a proper compact element h is big (ad hoc name) if h is below only one maximal element M(h). We separate in two cases.

Case (1): Every proper compact element is below some big compact element. We proceed by induction. Suppose that \(n\ge 1\) and we already have big compacts \(a_1\),...,\(a_n\) such that \(M(a_1)\),...,\(M(a_n)\) are distinct maximal elements (for \(n=1\) just pick an arbitrary big compact \(a_1\)). Let p be a maximal element other than \(M(a_1)\),...,\(M(a_n)\). There exists a compact element \(c\le p\) such that \(c\not \le M(a_i)\) for \(1\le i\le n\) (take \(c=c_1\vee \cdots \vee c_n\) where each \(c_i\in L^*\) satisfies \(c_i\le p\) and \(c_i\not \le M(a_i)\)). Then take a big compact element \(a_{n+1}\ge c\). This way we construct an infinite set \(\{ a_n;\ n\ge 1\}\) of big compacts such that all \(M(a_n)\)’s are distinct. Hence the \(a_n\)’s are pairwise comaximal.

Case (2): There exists a proper compact element \(a_0\) which is not below any big compact element. Clearly every proper compact above \(a_0\) inherits this property. Pick two distinct maximal elements p and q above \(a_0\). As L is a C-lattice, there exist two comaximal compacts \(a_1\le p\) and \(b_1\le q\) (note that \(p\vee q=1\), express p and q as joins of compact elements and use the fact that 1 is compact). Repeating this argument for \(a_1\), there exist two comaximal proper compacts \(a_2\ge a_1\) and \(b_2\ge a_1\). Note that \(b_2\) and \(b_1\) are comaximal. Thus we construct inductively an infinite set \(\{b_n ;\ n\ge 1\}\) of pairwise comaximal proper compact elements. \(\square \)

In a C-lattice L, we say that an element x is locally compact if \(x_m\) is compact in \(L_m\) for each \(m\in Max(L)\). We state our main result which is a Bazzoni-type theorem for C-lattices.

Theorem 5

Let L be a C-lattice domain satisfying the following two conditions:

(a) every nonzero element is above some cancellative compact element, and

(b) every compact element \(x\ne 1\) has some power \(x^n\) below some proper CMP element.

Then the following conditions are equivalent:

(i) Every locally compact element of L is compact.

(ii) Every nonzero element is below at most finitely many maximal elements.

Proof

\((ii) \Rightarrow (i).\) Although this part is well-known and easy, we include a proof for the reader’s convenience. Let x be a nonzero locally compact element of L and let \(a\le x\) be a nonzero compact element. By (ii), there are only finitely many maximal elements above a, say \(m_1\),...,\(m_k\). For each i between 1 and k, pick a compact element \(c_i\le x\) such that \(x_{m_i}=({c_i})_{m_i}\). A local check shows that \(x=a\vee c_1 \vee \cdots \vee c_k\), so x is compact. Note that this part works for any C-lattice.

\((i) \Rightarrow (ii)\). Deny, so suppose that some nonzero element c is below infinitely many maximal elements. By hypothesis (a), we may assume that c is a cancellative compact element. By Lemma 4 and hypothesis (b), there exist an infinite set \(\{ b_n;\ n\ge 1\}\) of proper pairwise comaximal CMP elements and integers \(k_n\ge 1\) such that \(c^{k_n}\le b_n\) for \(n\ge 1\) (\(k_n\) minimal with this property). Restricting to a subsequence, we may assume that \(k_n\le k_{n+1}\) for all n. We then have \(c^{k_n}\le b_1\wedge \cdots \wedge b_n=b_1\cdots b_n\) for all n.

Claim \((*):\) The element \(a:=\bigvee _{n\ge 1} (c^{k_n}:b_1\cdots b_n)\) is locally compact.

Pick \(m\in \) Max(L). Since the \(b_n\)’s are pairwise comaximal, m is above at most one of them. Assume first that \(m\ge b_s\). Since each product \( b_1\cdots b_n\) is compact, we get

$$a_m=(\bigvee _{n\ge 1} ((c^{k_n})_m:(b_1\cdots b_n)_m))_m= (c^{k_1}\vee (c^{k_s}:b_{s}))_m$$

which is compact in \(L_m\), cf. Lemma 3. Similarly, when m is above no \(b_n\), we get \(a_m=(c^{k_1})_m\), so \(a_m\) is compact in \(L_m\), hence Claim \((*)\) is proved. By (i), a is compact. So \(a=\bigvee _{n= 1}^q (c^{k_n}:b_1\cdots b_n)\) for some \(q\ge 1\). We get

$$ (c^{k_{q+1}}:b_1\cdots b_{q+1}) \le (c^{k_1}:b_1\cdots b_{q}) $$

so multiplying by \(b_1\cdots b_{q+1}\) (which is a CMP element) and taking into account that \(c^{k_{q+1}}\le b_1\cdots b_{q+1}\), we get

$$ c^{k_{q+1}} \le (c^{k_1}:b_1\cdots b_{q})b_1\cdots b_{q+1}\le c^{k_1}b_{q+1}. $$

Since \({k_{q+1}} \ge k_1\) and \(c^{k_1}\) is cancellative, we get \(c^{k_{q+1}-k_1}\le b_{q+1}\), which is a contradiction since \(k_{q+1}\) was minimal with \(c^{k_{q+1}}\le b_{q+1}\). \(\square \)

Recall that a C-lattice domain is a Prüfer lattice if every compact element is principal (i.e., meet-principal and join-principal). In a C-lattice domain, every nonzero join-principal element x is cancellative (because \((yx:x)=y\vee (0:x)=y\) for each y). So in a Prüfer lattice domain every nonzero compact element is CMP.

Corollary 6

Let L be a C-lattice domain in which every nonzero compact element is CMP (e.g., a Prüfer lattice domain). Then conditions (i) and (ii) of Theorem 5 are equivalent.

Bazzoni’s conjecture for Prüfer domains [10, Theorem 10] (see Introduction) follows from Corollary 6 since the ideal lattice of a Prüfer domain is clearly a Prüfer lattice.

3 Almost Prüfer Extensions

We recall several definitions from [6, 12]. Let \(A\subseteq B\) be a commutative ring extension and I an ideal of A. Then I is called B -regular if \(IB=B\) and I is called B -invertible if \(IJ=A\) for some A-submodule J of B. Every B-invertible ideal is B-regular, since \(A=IJ\subseteq IB\) implies \(IB=B\). We say that \(A \subseteq B\) is an almost Prüfer extension if every finitely generated B-regular ideal of A is B-invertible.

Finocchiaro and Tartarone [6, Theorem 4.5] proved Bazzoni’s conjecture for almost Prüfer ring extensions. We state their result and derive it from Corollary 6.

Theorem 7

(Finocchiaro and Tartarone) If \(A\subseteq B\) is an almost Prüfer extension, the following are equivalent:

(i) Every B-regular locally principal ideal of A is B-invertible.

(ii) Every B-regular ideal of A is contained in only finitely many maximal ideals of A.

Proof

It is well-known and easy to prove that (ii) implies (i), see [6, Corollary 3.5]. We prove the converse. Let L be the set of all B-regular ideals of A together with the zero ideal and order L by inclusion. As shown in [15, Lemma 7.1], L is a C-lattice domain under usual ideal multiplication, where the join is the ideal sum and the meet is the ideal intersection except the case when we get a non-B-regular ideal when we put \(\bigwedge =0\). By [15, Lemma 7.1], the set \(L^*\) of compact elements in L is exactly the set of (B-regular) finitely generated ideals of A together with the zero ideal. After this preparation it becomes clear that \([(i)\Rightarrow (ii)]\) follows from Corollary 6 provided we prove the two claims below. Write \(x\in L\) as \(\widehat{x}\) when considered as an ideal of A.

Claim 1: Every nonzero compact element of L is a CMP element.

Let c be a nonzero compact element of L. As \(A\subseteq B\) is almost Prüfer, \(\widehat{c}\) is a B-invertible ideal, so \(\widehat{c}J=A\) for some A-submodule J of B. Then c is clearly cancellative. By Lemma 3, it suffices to show that \((x:c)c=x\wedge c\) for each \(x\in L\). Changing x by \(x\wedge c\), we may assume that \(x\le c\). We have \(\widehat{x}=\widehat{x}J\widehat{c}\), so \(x=yc\) where \(y\in L\) is such that \(\widehat{y}=\widehat{x}J\) (note that \(\widehat{x}J\subseteq A\)). From \(x=yc\) we get \(y\le (x:c)\), so \(x=yc\le (x:c)c\le x\), thus \((x:c)c=x\).

Claim 2: Every locally compact element of L is compact.

Suppose that c is a nonzero locally compact element of L. Let m be a maximal element of L, that is, \(\widehat{m}\) is a B-regular maximal ideal of A. So \(c_m=\bigvee \{y\in L^*;\ ys\le c\) for some \(s\in L^*,s\not \le m\}\) is compact in the lattice \(L_m=\{x_m;\ x\in L\}\). Then \(c_m=h_m\) for some \(h\in L^*\). Extending these ideals in \(A_{\widehat{m}}\), we get \(\widehat{c}A_{\widehat{m}} = \widehat{c_m}A_{\widehat{m}}= \widehat{h_m}A_{\widehat{m}} = \widehat{h}A_{\widehat{m}}\). Since \(A\subseteq B\) is almost Prüfer, \(\widehat{h}\) is B-invertible. By [12, Proposition 2.3], \(\widehat{h}A_{\widehat{m}} = \widehat{c}A_{\widehat{m}}\) is a principal ideal of \(A_{\widehat{m}}\). Thus \(\widehat{c}\) is a locally principal ideal of A. By (i), \(\widehat{c}\) is B-invertible, so \(\widehat{c}\) is finitely generated, cf. [12, Proposition 2.3]. Thus c is compact in L. \(\square \)

4 Ideal Systems on Monoids and Integral Domains

Let H be a commutative multiplicative monoid (with zero element 0 and unit element 1) such that every nonzero element of H is cancellative and let \(\mathcal {P}(H)\) be the power set of H. A map \(r:\mathcal {P}(H)\rightarrow \mathcal {P}(H)\), \(X\mapsto X_r\), is an ideal system on H if the following conditions hold for all \(X,Y\in \mathcal {P}(H)\) and \(c\in H\):

$$(1)\ cX_r = (cX)_r, (2)\ X\subseteq X_r, (3)\ X\subseteq Y \text{ implies } X_r\subseteq Y_r, (4)\ (X_r)_r=X_r.$$

Then \(X_r\) is called the r -closure of X and a set of the form \(X_r\) is called an r-ideal. An r-ideal I is r -finite if \(I=Y_r\) for some finite subset Y of I. The ideal system r is called finitary if \(X_r=\bigcup \{ Y_r;\ Y\subseteq X\) finite\(\}\).

Assume that r is finitary. A proper r-ideal P is prime if, for \(x,y\in H\), \(xy\in P\) implies \(x\in P\) or \(y\in P\). The localization of H at P is the fraction monoid \(H_P=\{ x/t;\ x\in H, t\in H-P\}\) which comes together with the finitary ideal system \(r_P\) defined by \(\{a_1/s_1,...,a_n/s_n\}_{r_P}=(\{a_1,...,a_n\}_r)_P\) for all \(a_1,...,a_n\in H\) and \(s_1,...,s_n\in H-P\). A maximal r -ideal is a maximal element of the set of proper r-ideals of H. Any proper r-ideal is contained in a maximal one and a maximal r-ideal is prime. A nonzero r-ideal I is r -invertible if I is r-finite and r -locally principal (i.e., \(I_M=\{ x/t;\ x\in I, t\in H-M\}=yH_M\) with \(y\in I_M\) (depending on M) for all maximal r-ideals M). Next H is an r -Prüfer monoid if every nonzero r-finite r-ideal of H is r-invertible. For complete details we refer to [8].

Halter-Koch [9, Theorem 6.11] proved Bazzoni’s conjecture for r-Prüfer monoids. We state his result and derive it from Corollary 6.

Theorem 8

(Halter-Koch) If H is an r-Prüfer monoid for some finitary ideal system r on H, the following are equivalent.

(i) Every r-locally principal r-ideal of H is r-finite.

(ii) Every nonzero r-ideal of H is contained in only finitely many maximal r-ideals.

Proof

It is well-known and easy to prove that (ii) implies (i). We prove the converse. Let L be the set of all r-ideals of H ordered by inclusion. As shown in [8, Chapter 8], L is a C-lattice domain under r-ideal multiplication \((I,J)\mapsto (IJ)_r\), join \(\bigvee \{ J_\alpha \}:=(\bigcup J_\alpha )_r\) and meet \(\bigwedge \{ J_\alpha \}:=\bigcap J_\alpha \). By [15, Lemma 8.1], the set \(L^*\) of compact elements in L is exactly the set of r-finite r-ideals of H. After this preparation it becomes clear that \([(i)\Rightarrow (ii)]\) follows from Corollary 6 provided we prove the two claims below. Write \(x\in L\) as \(\widehat{x}\) when considered as an r-ideal of H.

Claim 1: Every nonzero compact element of L is a CMP element.

Let \(c\in L\) be a nonzero compact, in other words \(\widehat{c}\) is nonzero r-finite r-ideal. Since H is r-Prüfer, \(\widehat{c}\) is r-invertible. So c is a CMP element of L, cf. [15, Lemma 8.2].

Claim 2: Every locally compact element of L is compact.

Suppose that c is a locally compact element of L. Let m be a maximal element of L, that is, \(\widehat{m}\) is a maximal r-ideal of H. So \(c_m=\bigvee \{y\in L^*;\ ys\le c\) for some \(s\in L^*,s\not \le m\}\) is compact in the lattice \(L_m=\{x_m;\ x\in L\}\). Then \(c_m=h_m\) for some \(h\in L^*\). Switching to H, we have \((\widehat{c})_{\widehat{m}}=(\widehat{c_m})_{\widehat{m}}= (\widehat{h_m})_{\widehat{m}}=\widehat{h}_{\widehat{m}}\). Since H is an r-Prüfer monoid, \(\widehat{h}_{\widehat{m}}\) is a principal \(r_{\widehat{m}}\)-ideal of \(H_{\widehat{m}}\), cf. [8, Theorem 12.3]. Thus \(\widehat{c}\) is an r-locally principal r-ideal. By (i), \(\widehat{c}\) is r-finite, thus c is compact in L. \(\square \)

Next we present an application for integral domains. Let D be an integral domain. The t ideal system on D is defined by \(X_t=\bigcup \{ Y_v;\ Y\subseteq X \text{ finite }\}\) for all \(X\subseteq D\), where \(Y_v=\bigcap \{ (aD:_D b);\ a,b\in D, bY\subseteq aD\}.\) Clearly t is finitary, so the set \(Max_t(D)\) of maximal t-ideals is nonempty, see [8, Chapter 11].

The w ideal system on D is defined by \(X_w=\bigcap \{ XD_M;\ M\in Max_t(D)\}\) for all \(X\subseteq D\), where \(XD_M\) is the ideal generated by X in \(D_M\). So a w-ideal is an ideal of the ring D. For \(X\subseteq D\), we have \(X_w=(XD)_w\) and \(X_wD_M=XD_M\) for each \(M\in Max_t(D)\). Moreover, w is finitary and the set of maximal w-ideals is exactly \(Max_t(D)\). A w-finite ideal has the form \(((a_1,...,a_n)D)_w\) for some \(a_i\)’s in D. And a nonzero w-ideal is w -invertible if it is w-finite and t-locally principal (i.e., \(ID_M\) is a principal ideal of \(D_M\) for each \(M\in Max_t(D))\). For details on the w ideal system we refer to [2, 5].

According to [13], D is an almost Prüfer v -multiplication domain (in short APVMD) if for every \(a_1,...,a_n\in D-\{0\}\), the ideal \(((a^k_1,...,a^k_n)D)_w\) is w-invertible for some \(k\ge 1\). Say that a w-ideal I of D is t -locally finitely generated, if \(ID_M\) is a finitely generated ideal of \(D_M\) for each \(M\in Max_t(D)\).

Chang and Hamdi [4, Theorem 2.4] proved Bazzoni’s conjecture for APVMDs. We state their result and derive it from Theorem 5.

Theorem 9

(Chang and Hamdi) For an APVMD D, the following statements are equivalent:

(i) Each nonzero t-locally finitely generated w-ideal of D is w-finite.

(ii) Every nonzero ideal of D is contained in only finitely many maximal t-ideals.

Proof

It is well-known and easy to prove that (ii) implies (i), see for instance the proof of \((3)\Rightarrow (1)\) in [4, Theorem 2.4]. We prove that (i) implies (ii). Let L be the set of all w-ideals of D ordered by inclusion. As shown in [8, Chapter 8], L is a C-lattice domain under w-ideal multiplication \((I,J)\mapsto (IJ)_w\), join \(\bigvee \{ J_\alpha \}:=(\bigcup J_\alpha )_w\) and meet \(\bigwedge \{ J_\alpha \}:=\bigcap J_\alpha \). By [15, Lemma 8.1], the set \(L^*\) of compact elements in L is exactly the set of w-finite ideals of D. Condition (a) of Theorem 5 holds clearly for L (any nonzero ideal contains a nonzero principal ideal). After this preparation it becomes clear that \([(i)\Rightarrow (ii)]\) follows from Theorem 5 provided we prove the two claims below. Write \(x\in L\) as \(\widehat{x}\) when considered as a w-ideal of D.

Claim 1: Every \(d\in L^*-\{1\}\) has some power \(d^n\) below some proper CMP element.

Let \(d\in L^*-\{1\}\). Then \(\widehat{d}=(a_1,...,a_k)_w\) for some elements \(a_i\in \widehat{d}\). Since D is an APVMD, \((a^s_1,...,a^s_k)_w\) is w-invertible for some \(s\ge 1\). If \((a^s_1,...,a^s_k)_w=\widehat{f}\) with \(f\in L\), then f is a proper CMP element of L, cf. [15, Lemma 8.2]. Moreover \(d^{sk}\le f\), so Claim 1 is proved.

Claim 2: Every locally compact element of L is compact.

Suppose that c is a locally compact element of L. Let m be a maximal element of L, that is, \(\widehat{m}\) is a maximal w-ideal of D. So \(c_m=\bigvee \{y\in L^*;\ ys\le c\) for some \(s\in L^*,s\not \le m\}\) is compact in the lattice \(L_m=\{x_m;\ x\in L\}\). Then \(c_m=h_m\) for some \(h\in L^*\). Hence \(\widehat{h}=(b_1,...,b_n)_w\) for some elements \(b_i\in \widehat{h}\). Switching to D, we have

$$\widehat{c}D_{\widehat{m}}=\widehat{c_m}D_{\widehat{m}}= (\widehat{h_m})D_{\widehat{m}}=\widehat{h}D_{\widehat{m}} =(b_1,...,b_n)_wD_{\widehat{m}}=(b_1,...,b_n)D_{\widehat{m}}.$$

Thus \(\widehat{c}\) is a t-locally finitely generated ideal of D. By (i), \(\widehat{c}\) is w-finite, thus c is compact in L. \(\square \)