1 Introduction

A notion of derivation whose origin is in analysis has been applied to theory of algebras with two operations \(+\) and \(\cdot \), especially to the theory of rings (Posner 1957). For an algebra \(A=(A,+,\cdot )\), a map \(f:A\rightarrow A\) is called a derivation in Posner (1957) if it satisfies the conditions: For all \(x,y\in A\),

$$\begin{aligned}&f(x+y) = f(x) + f(y) \\&f(x\cdot y) = f(x) \cdot y + x\cdot f(y). \end{aligned}$$

It was applied to the theory of lattices (Ferrari 2001; Szász 1975; Xin et al. 2008), where operations \(+\) and \(\cdot \) were interpreted as lattice operations \(\vee \) and \(\wedge \), respectively. In particular, it has been proven in Szász (1975) that if f is a derivation of a bounded lattice L, that is, f satisfies two conditions:

$$\begin{aligned} f(x\vee y)&= f(x) \vee f(y) \\ f(x\wedge y)&= (f(x) \wedge y) \vee (x\wedge f(y)), \end{aligned}$$

then f just has a form \(f(x) = x \wedge f(1)\). Since then, every type of derivations is defined as a map \(f: X \rightarrow X\) satisfying only a condition:

$$\begin{aligned} f(x\cdot y) = f(x) \cdot y + x\cdot f(y), \end{aligned}$$

where \((X,+, \cdot )\) is an algebra. Thus, in the case of lattices, a derivation f of a lattice L is defined a map \(f:L \rightarrow L\) satisfying the condition:

$$\begin{aligned} f(x\wedge y) = (f(x) \wedge y) \vee (x\wedge f(y)). \end{aligned}$$

By the use of derivations of lattices, characterization theorems of distributive lattices and of modular lattices were proven in [6]:

Let L be a lattice and f a derivation.

  1. (1)

    The condition that f is monotone \(\Leftrightarrow f(f(x)\vee y) = f(x) \vee f(y)\ (\forall x,y\in L)\) holds if and only if L is a modular lattice.

  2. (2)

    The condition that f is monotone \(\Leftrightarrow f(x\vee y) = f(x) \vee f(y) \ (\forall x,y\in L)\) holds if and only if L is a distributive lattice.

Further, it is also applied to other algebras, such as MV algebras (Alshehri 2010; Ghorbani et al. 2013; Yazarli 2013), where operations \(+\) and \(\cdot \) were interpreted as \(\oplus \) and \(\odot \), respectively. In He et al. (2016), another type of derivations, multiplicative derivations, are defined on residuated lattices, where operations \(+\) and \(\cdot \) are interpreted as \(\vee \) and \(\odot \), respectively. By this definition of derivations, we have several interesting properties about residuated lattices. For example, in He et al. (2016), it is shown that if d is an ideal derivation of a residuated lattice L, then the set \(\hbox {Fix}_d(L)\) of all fixed elements formed a residuated lattice. Thus, it is expectable for residuated lattices with multiplicative derivations to have other deeper properties. Indeed, we show some interesting properties here, for example, for every residuated lattice L with an ideal multiplicative derivation d, the set \(\hbox {Fix}_d(L)\) is isomorphic to \(L/\ker (d)\) and hence \(\hbox {Fix}_d(L)\) is a residuated lattice.

In this paper, we show the following results. Let L be a commutative residuated lattice and d be an ideal derivation of it. Then, we have

  1. (1)

    The set \(\hbox {Fix}_d(L)\) of all fixed points of d forms a residuated lattice and \(L/{\ker (d)} \cong \hbox {Fix}_d(L)\);

  2. (2)

    A map \(d/F: L/F \rightarrow L/F\) defined by \((d/F)(x/F) = dx/F\) is also an ideal derivation of L/F;

  3. (3)

    The quotient residuated lattices \(\hbox {Fix}_{d/F}(L/F)\) and \(\hbox {Fix}_d(L)/d(F)\) are isomorphic, namely \(\hbox {Fix}_{d/F}(L/F) \cong \hbox {Fix}_d(L)/d(F)\).

We also show the characterization theorem of d-filters, which says that for an ideal derivation d and a non-empty subset S, the smallest d-filter containing S is identical with the filter containing \(S\cup d(S)\), that is, \([S)_d = [S\cup d(S))\).

2 Derivations of residuated lattices

We recall a definition of bounded integral commutative residuated lattices (Galatos et al. 2007; Ward and Dilworth 1939). An algebraic structure \((L, \wedge , \vee , \odot , \rightarrow , 0, 1)\) is called a bounded integral commutative residuated lattice (simply called a residuated lattice) if

  1. (1)

    \((L, \wedge , \vee , 0, 1)\) is a bounded lattice;

  2. (2)

    \((L, \odot , 1)\) is a commutative monoid with unit element 1;

  3. (3)

    For all \(x, y, z \in L\), \(x \odot y \le z\) if and only if \(x \le y \rightarrow z\).

For all \(x \in L\), by \(x'\), we mean \(x' = x \rightarrow 0\), which is a negation in a sense.

In what follows, let \(L=(L,\wedge , \vee , \odot , \rightarrow ,0,1)\) be a residuated lattice. An element \(x\in L\) is called complemented if there exists an element \(y\in L\) such that \(x \wedge y=0\) and \(x\vee y=1\). By B(L), we mean the set of all complemented elements of L, i.e.,

$$\begin{aligned} B(L) = \{ x\in L \,|\, \exists y\in L \ \mathrm{s.t.} \ x\wedge y=0, x\vee y=1 \}. \end{aligned}$$

It is easy to show the following results, so we omit their proofs.

Proposition 1

(Galatos et al. 2007) For a residuated lattice L, we have

  1. (1)

    \(x\in B(L)\) if and only if \(x\vee x'=1\);

  2. (2)

    If \(x\in B(L)\), then \(x\wedge y= x\odot y\) for all \(y\in L\);

  3. (3)

    If \(x\oplus y= x\vee y\) for all \(y\in L\), then \(x\vee x'=1\), where \(x\oplus y = (x'\odot y')'\);

  4. (4)

    B(L) is a Boolean subalgebra of L.

We note that \(a\odot (a\rightarrow x)=a\odot x\) for all \(a\in B(L)\) and \(x\in L\), because, since \(a\in B(L)\) and \(a\odot (a\rightarrow x)\le x\), we have \(a\odot (a\rightarrow x)=a\odot a\odot (a\rightarrow x) \le a\odot x\). Conversely, it follows from \(x\le a\rightarrow x\) that \(a\odot x\le a\odot (a\rightarrow x)\). Therefore, we get \(a\odot (a\rightarrow x)=a\odot x\) for all \(a\in B(L)\) and \(x\in L\).

We have the following basic properties of residuated lattices (Galatos et al. 2007).

Proposition 2

For all \(x, y, z \in L\), we have

  1. (1)

    \(0'= 1, 1'= 0\);

  2. (2)

    \(x \odot x'= 0\);

  3. (3)

    \(x \le y \ \Longleftrightarrow \ x \rightarrow y = 1\);

  4. (4)

    \(x \odot (x \rightarrow y) \le y\);

  5. (5)

    \(x \le y \ \Longrightarrow \ x \odot z \le y \odot z, z \rightarrow x \le z \rightarrow y, y \rightarrow z \le x \rightarrow z\);

  6. (6)

    \(1 \rightarrow x = x\);

  7. (7)

    \((x \vee y) \odot z = (x \odot z) \vee (y \odot z)\);

  8. (8)

    \((x \vee y)' = x' \wedge y'\);

  9. (9)

    \((x''\odot y'')^{''} =(x\odot y'')^{''}= (x\odot y)^{''}\).

We define derivations of residuated lattices according to He et al. (2016). A map \(d: L \rightarrow L\) is called a multiplicative derivation (or simply derivation here) of L if it satisfies the condition:

$$\begin{aligned} d(x\wedge y) = (dx \odot y) \vee (x \odot dy) \ \ \ (\forall x,y\in L). \end{aligned}$$

We simply denote dx instead of d(x). A derivation d is called monotone; if \(x\le y\), then \(dx\le dy\), that is, d is order-preserving. A derivation d is good when \(d1\in B(L)\). If a derivation d is monotone and good, then it is said to be ideal. We note that any multiplicative derivation d is contractive, \(dx\le x\) for all \(x\in L\), because we have \(dx = d(x\wedge x) = (dx\odot x) \vee (x\odot dx) = dx\odot x \le x\). This result was not referred in He et al. (2016).

Example

Let \(X=\{0,a,1\}\) with \(0<a<1\) be a residuated lattice if we define \(x\wedge y=x\odot y=\min \{x,y\}, x\vee y=\max \{x,y\}\) and

$$\begin{aligned} x\rightarrow y=\left\{ \begin{array}{ll} 1 &{}\quad (\mathrm{if} \ x\le y) \\ y &{}\quad (\mathrm{otherwise}) \\ \end{array}. \right. \end{aligned}$$

A map \(d_a\) defined by \(d_a(x)=x\wedge a\) for all \(x\in X\) is a monotone derivation, but not good. As another example of derivation, we have \(f:X\rightarrow X\) defined by \(f1=0=f0, fa=a\). It is easy to show that f is a good derivation, but it is not monotone.

We have fundamental results about derivations of residuated lattices.

Proposition 3

(He et al. 2016) Let d be a derivation of L. For all \(x,y\in L\),

  1. (1)

    \(d0=0\);

  2. (2)

    \(dx \ge x\odot d1\);

  3. (3)

    \(dx^n = x^{n-1}\odot dx\) for all \(n \ge 1\);

  4. (4)

    If \(x\odot y=0\) then \(dx \odot y = x\odot dy = dx \odot dy =0\);

  5. (5)

    \(d(x') \le (dx)'\).

In He et al. (2016), the following characterization theorem about an ideal derivation was proven.

Theorem 1

(He et al. 2016) Let d be a derivation of L and \(d1\in B(L)\). Then, the following are equivalent: for all \(x,y\in L\),

  1. (1)

    d is an ideal derivation;

  2. (2)

    \(dx\le d1\);

  3. (3)

    \(dx= x\odot d1= x\wedge d1\);

  4. (4)

    \(d(x\wedge y) = dx \wedge dy\);

  5. (5)

    \(d(x\vee y) = dx\vee dy\);

  6. (6)

    \(d(x\odot y) = dx \odot dy\).

From the above, we see that \(dx=d(dx)\) for all \(x\in L\) for an ideal derivation d, that is, \(d=d^2\), because, since d is the ideal derivation, we have \(dx= x\wedge d1\) and thus \(d^2x = d(dx) =d(x\wedge d1) = (x\wedge d1)\wedge d1 = x\wedge d1 = dx\) for all \(x\in L\). This means that \(d^2 = d\).

For a derivation d of L, we denote by \(\hbox {Fix}_d(L)\) the set of all fixed elements of L for d, that is, \(\hbox {Fix}_d(L) =\{x\in L\,|\, dx=x \}\).

It is easy to show the next result; hence, we omit its proof.

Proposition 4

For an ideal derivation d of a residuated lattice L, we have \(\hbox {Fix}_d(L) = d(L)\).

Lemma 1

Let d be an ideal derivation. Then, we have \(d(dx\rightarrow dy) = d(x\rightarrow y)\) for all \(x,y\in L\).

Proof

Suppose that d is an ideal derivation. Since \(dz = z\odot d1=z\wedge d1\) for all \(z\in L\), we have

$$\begin{aligned} d(dx\rightarrow dy)&= (dx\rightarrow dy) \odot d1 \\&= (x\wedge d1 \rightarrow y\wedge d1)\odot d1 \\&= \{(x\wedge d1\rightarrow y) \wedge (x\wedge d1 \rightarrow d1)\} \odot d1 \\&= (x\wedge d1 \rightarrow y)\odot d1 \\&= d1\odot (d1\odot x \rightarrow y) \\&= d1 \odot (d1 \rightarrow (x\rightarrow y)) \\&= d1 \odot (x\rightarrow y) \ \ \ (d1 \in B(L))\\&= d(x\rightarrow y). \end{aligned}$$

\(\square \)

It follows from the characterization theorem about ideal derivations and above that we define some operations in \(\hbox {Fix}_d(L)=d(L)\) by

$$\begin{aligned} dx \sqcap dy&= d(x\wedge y); \\ dx\sqcup dy&= d(x\vee y); \\ dx\boxdot dy&= d(x\odot y); \\ dx \mapsto dy&= d(dx\rightarrow dy). \end{aligned}$$

Then, we have

Theorem 2

Let L be a residuated lattice and d be an ideal derivation of L. Then, \(\hbox {Fix}_d(L) = (\hbox {Fix}_d(L), \sqcap , \sqcup , \boxdot , \mapsto , 0, d1)\) is a residuated lattice. However, it is not a subalgebra of L in general.

Proof

We only show that \(dx \odot dy \le dz\) if and only if \(dx \le dy \mapsto dz\) for all \(dx, dy, dz\in d(L)=\hbox {Fix}_d(L)\). If \(dx\odot dy \le dz\), then we have \(dx\le dy \rightarrow dz\) and thus \(dx = d(dx)\le d(dy\rightarrow dz) = dy \mapsto dz\). Conversely, suppose that \(dx\le dy\mapsto dz=d(dy\rightarrow dz)\). Since d is contractive, we have \(d(dy\rightarrow dz)\le dy\rightarrow dz\) and hence \(dx\le dy\rightarrow dz\). This yields \(dx\odot dy\le dz\). \(\square \)

Remark 1

We note that the theorem above was already proven as Theorem 3.15 in He et al. (2016), where the meet operation \(\sqcap \) is defined by \(d(dx\wedge dy)\).

We have proved above that \(\hbox {Fix}_d(L)\) is a residuated lattice for an ideal derivation d. Moreover, we see that any ideal derivation d is a homomorphism from L to \(\hbox {Fix}_d(L)\), since \(d(x\rightarrow y) = d(dx\rightarrow dy) = dx\mapsto dy\). The other cases such as \(d(x\wedge y)\) also can be proved easily. Therefore, d is the homomorphism from L to \(\hbox {Fix}_d(L)\). It follows from the homomorphism theorem of residuated lattices that \(L/\ker (d) \cong \hbox {Fix}_d(L)\), where \(\ker (d) = \{(x,y) \,|\, dx=dy \}\).

Theorem 3

For every ideal derivation d, it is a homomorphism from L to \(\hbox {Fix}_d(L)\) and hence

$$\begin{aligned} L/\ker (d) \cong \hbox {Fix}_d(L). \end{aligned}$$

3 Galois connection of derivations

In this section, we consider Galois connections of derivations. Let PQ be partially ordered sets and fg be maps, \(f:P \rightarrow Q\), \(g: Q \rightarrow P\). A pair (fg) of maps is called a Galois connection if

$$\begin{aligned} f(x)\le _{Q} y \ \Leftrightarrow \ x\le _{P} g(y) \ \ (\forall x\in P, \forall y\in Q). \end{aligned}$$

A Galois connection (fg) is simply denoted by \(f \dashv g\).

Let L be a residuated lattice and d be an ideal derivation of L. By the characterization theorem of ideal derivations, d has the form \(dx=x\odot d1 = x\wedge d1\). In this case, we may ask

$$\begin{aligned} \text {``Is there a map}\ g\ \hbox {such that}\ d \dashv g?\text {''} \end{aligned}$$

We have an affirmative solution as follows.

Theorem 4

Let d be an ideal derivation. There exists an ideal derivation \(g: L\rightarrow L\) such that \(d \dashv g\). Moreover, g is idempotent.

Proof

We define \(gx = d1 \rightarrow x\) for all \(x\in L\). It is obvious that \(dx=x\odot d1 \le y\) if and only if \(x\le d1 \rightarrow y= gy\), that is, \(dx\le y \ \Leftrightarrow x\le gy\) for all \(x,y\in L\). Moreover, since \(g(gx)=d1 \rightarrow (d1 \rightarrow x) = d1 \odot d1 \rightarrow x = d1 \rightarrow x = gx\), g is idempotent. \(\square \)

From the general theory of idempotent Galois connection, that is, \(f\dashv g\), \(f^2 =f\) and \(g^2=g\) that two subsets \(F_f= \{x\in L \,|\, fx=x\}\) and \(F_g= \{x\in L \,|\, gx=x\}\) are isomorphic as partially ordered sets. Hence, we have the following result.

Theorem 5

For an ideal derivation d, \(\hbox {Fix}_{d}(L) =F_{d} (L) \cong F_{g}(L) = \hbox {Fix}_{g} (L)\) as partially ordered sets.

4 d-filter and its characterization

We define a filter of a residuated lattice which plays an important role in this paper. Let F be a non-empty subset of L. We call F a filter of L if it satisfies the following conditions: For all \(x,y\in L\),

  1. (F1)

    if \(x,y\in F\), then \(x\odot y\in F\);

  2. (F2)

    if \(x\in F\) and \(x\le y\), then \(y\in F\).

Let d be an ideal derivation of L. A filter F of L is called an ideal derivation filter (simply d-filter here) if \(x\in F\) implies \(dx\in F\) for all \(x\in L\). By \({\mathcal {F}}(L)\) (or \({{\mathcal {F}}}_d(L)\)), we mean the set of all filters (or d-filters) of L. By [S) (or \([S)_d\)), we mean the generated filter (or generated d-filter, respectively) by S. At first, we provide a characterization theorem about d-filters.

Theorem 6

Let d be an ideal derivation of L. For a non-empty subset S of L, we have \([S)_d = [S\cup d(S))\).

Proof

It is sufficient to show that \([S\cup d(S))\) is the least d-filter including S. It is obvious \(S \subseteq [S\cup d(S))\). We show that \([S\cup d(S))\) is a d-filter, that is, if \(x\in [S\cup d(S))\), then \(dx\in [S\cup d(S))\). Suppose that \(x\in [S\cup d(S))\). There exist \(a_i, b_j\in S\) such that \(a_1 \odot \cdots \odot a_m \odot db_1 \odot \cdots db_n \le x\). Since d is the ideal derivation, we have \(da_1 \odot \cdots \odot da_m \odot db_1 \odot \cdots db_n =da_1 \odot \cdots \odot da_m \odot d(db_1) \odot \cdots d(db_n) =d(a_1 \odot \cdots \odot a_m \odot db_1 \odot \cdots db_n) \le dx\). It follows from \(da_i, db_j \in d(S) \subseteq S \cup d(S)\) that \(dx \in [S\cup d(S))\). Therefore, \([S\cup d(S))\) is the d-filter. Lastly, for any d-filter F including S, we have \(d(S) \subseteq d(F) \subseteq F\) and hence \(S\cup d(S) \subseteq F\). This implies \([S\cup d(S)) \subseteq F\). Therefore, \([S\cup d(S))\) is the least d-filter containing S and thus \([S)_d = [S\cup d(S))\). \(\square \)

Corollary 1

(Theorem 3.19 He et al. 2016) The class \({\mathcal {F}}_d(L)\) of all d-filters forms a complete Heyting subalgebra of the class \({\mathcal {F}}(L)\) of all filters of a residuated lattice L

Corollary 2

If F is a filter, then \([F)_d = [d(F))\).

Corollary 3

For any \(F\in {\mathcal {F}}(L)\), F is a d-filter if and only if \(F=[d(F))\).

Proof

It is easy to show that if F is a d-filter, then \(F=[F)_d=[d(F))\). Conversely, suppose that \(x\in F=[d(F))\). There exist \(a_i \in F\) such that \(da_1 \odot \cdots \odot da_n\le x\). If we take \(a= a_1 \odot \cdots \odot a_n \in F\), then \(da\in d(F)\). Moreover, we have \(da = d(a_1 \odot \cdots \odot a_n) = da_1 \odot \cdots \odot da_n \le x\) and thus \(da = d(da) \le dx\). This implies that \(dx\in [d(F))=F\), namely F is the d-filter. \(\square \)

Proposition 5

Let d be an ideal derivation of L. Then, we have

  1. (1)

    \(F\in {\mathcal {F}}_d(L) \Rightarrow d(F)\in {\mathcal {F}}(d(L))\);

  2. (2)

    \(G\in {\mathcal {F}}(d(L)) \Rightarrow d^{-1}(G)\in {\mathcal {F}}_d(L)\), where \(d^{-1}(G) = \{ x\in L\,|\, dx\in G\}\).

Proof

  1. (1)

    If \(dx, dy\in d(F), (x,y\in F)\) then, since \(x\odot y\in F\) and d is the ideal derivation, we have \(dx\odot dy = d(x\odot y) \in d(F)\). Moreover, if \(dx\in d(F)\) for \(x\in F\) and \(dx\le dy\), then \(dx\in F\) and thus \(dy\in F\). Then, \(dy = d(dy)\in d(F)\). Therefore, d(F) is a filter of d(L), that is, \(d(F)\in {\mathcal {F}}(d(L))\).

  2. (2)

    Let G be a filter of d(L). For all \(x,y\in d^{-1}(G)\), since \(dx, dy\in G\), we get \(dx\odot dy = d(x\odot y) \in G\) and \(x\odot y\in d^{-1}(G)\). Next, suppose that \(x\in d^{-1}(G)\) and \(x\le y\). It follows from \(dx\in G\) and \(dy\le dy\) that \(dy\in G\) and hence that \(y\in d^{-1}(G)\). Lastly, if \(x\in d^{-1}(G)\), then \(dx\in G\) and \(d(dx) = dx\in G\). This implies \(dx\in d^{-1}(G)\). Therefore, \(d^{-1}(G)\) is the d-filter, that is, \(d^{-1}(G)\in {\mathcal {F}}_d(L)\).\(\square \)

For a filter F, we define a quotient structure L/F, where \(x/F=y/F\) is defined by \(x\rightarrow y, y\rightarrow x\in F\) for all \(x,y\in L\). Since the set of all residuated lattices forms a variety, the quotient algebra L/F by a filter F is also a residuated lattice.

Proposition 6

Let d be an ideal derivation and F be a d-filter of L. If we define a map \(d/F : L/F \rightarrow L/F\) by \((d/F)(x/F) = dx/F\) for all \(x/F\in L/F\), then d/F is an ideal derivation of L/F.

Proof

At first, we show that d/F is well defined. Suppose that \(x/F=y/F\). Since \(x\rightarrow y, y\rightarrow x\in F(\in {\mathcal {F}}_d(L))\), we have \(d(x\rightarrow y), d(y\rightarrow x)\in F\). It follows from \(d(x\rightarrow y) \le dx\rightarrow dy, d(y\rightarrow x)\le dy\rightarrow dx\) that \(dx\rightarrow dy, dy\rightarrow dx\in F\). This means that \((d/F)(x/F)=dx/F=dy/F=(d/F)(y/F)\) and hence that d/F is well defined. It is easy to show that d/F is a good derivation of the residuated lattice L/F. It is sufficient to show that d/F is monotone. Suppose that \(x/F\le y/F\). Since \(x\rightarrow y\in F\) and F is the d-filter, we have \(d(x\rightarrow y)\in F\) and \(dx\rightarrow dy\in F\) because of \(d(x\rightarrow y)\le dx\rightarrow dy\). This implies \(dx/F \le dy/F\) and hence \((d/F)(x/F)=dx/F \le dy/F = (d/F)(y/F)\), namely d/F is monotone. \(\square \)

We note that \(d1/F=1/F\), because, since F is the d-filter, the fact \(1\in F\) implies \(d1\in F\), that is, \(d1/F=1/F\). It follows from the above that the quotient structure \((d/F)(L/F) =\hbox {Fix}_{d/F}(L/F)= (\hbox {Fix}_{d/F}(L/F), \wedge , \vee , \odot , \mapsto , 0/F, 1/F)\) is also a residuated lattice for any ideal derivation d and d-filter F. Moreover, since F is the d-filter, d(F) is the filter of d(L) and thus the quotient structure d(L)/d(F) forms a residuated lattice. It is natural to ask what the relation between two residuated lattices (d/F)(L/F) and d(L)/d(F) is. In order to answer the question, we need the following result.

Lemma 2

If \(F\in {\mathcal {F}}_d(L)\), then we have \(F \cap d(L) = d(F)\).

Proof

For \(x\in F \cap d(L)\), since \(x\in d(L)= \hbox {Fix}_d(L)\), we have \(x=dx\in d(F)\) and \(F \cap d(L) \subseteq d(F)\).

Conversely, suppose \(x\in d(F)\). There exists \(y\in F\) such that \(x=dy\). Since F is the d-filter, we get \(x=dy\in F\). On the other hand, it follows from \(dx = d(dy) = dy = x\) that \(x\in \hbox {Fix}_d(L)=d(L)\) and \(d(F)\subseteq F \cap d(L)\). Therefore, \(F \cap d(L) = d(F)\). \(\square \)

From the above, we have a following result which answers the question.

Theorem 7

Let d be an ideal derivation and F be a d-filter of L. Then, we have \((d/F)(L/F)= \hbox {Fix}_{d/F}(L/F)\) is isomorphic to d(L)/d(F), that is,

$$\begin{aligned} (d/F)(L/F)= \hbox {Fix}_{d/F}(L/F) \cong d(L)/d(F). \end{aligned}$$

Proof

We define a map \(\Phi : (d/F)(L/F) \rightarrow d(L)/d(F)\) by \(\Phi ((d/F)(x/F))=dx/d(F)\). We show that \(\Phi \) is an isomorphism. We only show that the map \(\Phi \) is well defined and injective. At first, \(\Phi \) is well defined, because we have

$$\begin{aligned} (d/F)(x/F)&=(d/F)(y/F) \\&\Rightarrow dx/F=dy/F \\&\Rightarrow dx\rightarrow dy, dy\rightarrow dx\in F \\&\Rightarrow d(dx\rightarrow dy), d(dy\rightarrow dx) \in d(F) \\&\Rightarrow dx\mapsto dy, dy\mapsto dx\in d(F) \\&\Rightarrow dx/d(F)=dy/d(F). \end{aligned}$$

Hence, \(\Phi \) is well defined.

Suppose \(\Phi ((d/F)(x/F))=\Phi ((d/F)(y/F))\). Since \(dx/d(F)=dy/d(F)\), we have \(dx\mapsto dy, dy\mapsto dx\in d(F)\) and thus \(d(dx\rightarrow dy), d(dy\rightarrow dx)\in d(F)= F\cap d(L)\). This implies \(d(dx\rightarrow dy), d(dy\rightarrow dx)\in F\). Since d is contractive, we also get \(d(dx\rightarrow dy)\le dx\rightarrow dy, d(dy\rightarrow dx)\le dy\rightarrow dx\) and \(dx\rightarrow dy, dy\rightarrow dx\in F\). Therefore, \(dx/F=dy/F\) and \((d/F)(x/F)= dx/F =dy/F = (d/F)(y/F)\), namely \(\Phi \) is injective. \(\square \)