1. Introduction
Recent advances in operator theory have highlighted the fundamental roles of idempotent operators across mathematical physics, where their norm properties and generalized inverses enable critical applications. Studies such as [
1] have revealed deep connections between algebraic equivalence, similarity, and norm behavior in nest algebras, while [
2] established geometric relationships between idempotent norms and subspace configurations in Hilbert spaces. Further investigations by [
3] led to the derivation of norm lower bounds for idempotent functions on locally compact groups, and [
4] uncovered concentration phenomena for integral norms of trigonometric polynomials. Constructive approaches in [
5] yielded idempotents with controlled diagonals, whereas [
6] developed explicit expressions for Drazin inverses of operator sums. Complementary work by [
7] allowed for the characterization of matrices with idempotent Moore–Penrose inverses, and [
8] quantified Frobenius distances to generalized inverses.
Quantum systems derive profound advantages from these developments, where idempotent norm constraints govern the interchangeability of projections and explain fundamental quantization phenomena such as integer quantum conductance in the Hall effect [
9]. Complementary perturbation analyses [
10] have demonstrated how the exponential decay properties of Fermi projections induce Anderson orthogonality with positive probability in localized systems. Beyond these phenomena, generalized inverses enable crucial advancements in quantum information processing, facilitating near-optimal reversal operations that are adaptive to noise and initial states while preserving the fidelity of both quantum entanglement and classical information for efficient state recovery in noisy channels [
11]. These techniques further extend to non-Hermitian quantum systems, where generalized inverses support the construction of metric-operator frameworks that ensure essential compatibility between probability conservation and real spectra in pseudo-Hermitian systems [
12]. Building upon these foundations, we establish the complete spectral characterization of conjugatable idempotent sums on Hilbert
-modules, determining both the Moore–Penrose inverse and spectral norm for the sum formed by a conjugatable idempotent operator and its adjoint.
Research on linear combinations of idempotents has yielded substantial insights, particularly through the complete characterization of idempotency for two-matrix combinations in [
13]. This foundation was extended to three-idempotent combinations with pairwise commutativity in [
14], while Drazin inverses for operator sums/differences emerged through the Hilbert space constraints presented in [
6] and the Banach algebra representations in [
15]. Despite these advances, decomposition theory for idempotents remains comparatively underdeveloped. The decomposition of certain idempotent matrices demonstrated by [
16], the proof in [
17] that every square matrix expresses as a three-idempotent combination, and the establishment that diagonal operators decompose into three-idempotent combinations (requiring four projections for self-adjoint cases) in [
18] collectively highlight a persistent gap: the unresolved characterization of two-term linear decompositions. This study addresses this theoretical deficiency within Hilbert
-module theory, providing comprehensive decomposition characterizations for arbitrary idempotents (including non-adjointable cases).
Factorization problems have also attracted considerable attention. Prior work has extensively studied operators that are expressible as products of projections [
19,
20,
21,
22,
23]. In [
24], the factorization of idempotents into products of two idempotents were examined, deriving explicit representations. In this paper, we extend these factorization results to the Hilbert
-module setting, significantly broadening their applicability.
The remainder of this paper is structured as follows:
Section 2 presents a generalized version of Halmos’ two-projections theorem for Hilbert
-modules and essential preliminaries. To enable the application of this generalized Halmos framework to idempotent operators,
Section 3 establishes necessary and sufficient conditions for the closedness and orthogonal complementability of ranges for sums of adjointable idempotents and their adjoints, supplemented by counterexamples demonstrating the failure of these properties in general cases.
Section 4 provides explicit characterizations of the Moore–Penrose inverse, spectral points, and a spectral norm for such sums.
Section 5 presents advances in two directions: (1) For adjointable idempotents on Hilbert
-modules, we derive concrete representations of linearly decomposed idempotents; and (2) we extend these results to non-adjointable idempotents, obtaining precise representations for both linear and multiplicative decompositions. Finally,
Section 6 presents: (i) Matrix decomposition examples illustrating computational applications of our theorems; and (ii) Hilbert
-module counterexamples validating the necessity of adjointability assumptions for range properties, Moore–Penrose inverses, and spectral norms.
2. Preliminaries
Throughout this paper,
denotes a
-algebra. We assume that
and
are Hilbert
-modules over
The set of all adjointable operators from
to
is represented by
, with the abbreviation
if
The identity element of an algebra is denoted by
The range and nullity of an operator
T are denoted by
and
, respectively. For further information on Hilbert
-modules and their geometry, see [
25,
26,
27].
An operator
is said to be idempotent if
. By a projection, we mean an operator
such that
. Recall that a submodule
is said to be orthogonally complemented in
if
, where
. In this case,
is closed, and we refer to the projection from
onto
as
. Unlike Hilbert spaces, a closed submodule is not necessarily orthogonally complemented. H In this paper, the notations “⊕" and “∔" are used with different meanings. For Hilbert
-modules
and
, let
which is also a Hilbert
-module whose
-valued inner product is given by
for any
Let
and
be submodules of a Hilbert
-module
. If
for all
and
, we define the orthogonal sum as follows:
This study utilizes Halmos’ two projections theorem as a mathematical tool. Halmos’ two projections theorem was originally obtained in [
28], more applications can be found in [
29,
30], and we generalized it to harmonious projection pairs on Hilbert
-modules in [
31]. A brief introduction is provided in the following.
Two projections
are said to be harmonious ([
31], Definition 4.1) if the four closures
are all orthogonally complemented in
H.
Suppose that
are two harmonious projections. Let
Since
and
is orthogonally complemented in
, we conclude that
is likewise orthogonally complemented in
. Similarly,
,
, and
are all orthogonally complemented in
.
Furthermore, let
and define
With the notation above, a unitary operator
can be induced as follows:
with the property that
It follows that
where
in which
is the restriction of the operator
on
. The same convention can be used for
and
.
Lemma 1 ([
31], Theorem 4.6).
Suppose that are two harmonious projections. Let ) be defined by (2)–(6), respectively. Then, the operator T formulated by (10) can be characterized aswhere is a unitary operator, is the restriction of on , and both and are positive, injective, and contractive. The following lemmas are crucial for the decomposition of idempotent operators, the closedness of operator ranges, and orthogonal complementability.
Lemma 2 ([
31], Lemma 2.3).
Let be two projections. Then, Lemma 3 ([
32], Theorem 3.2).
Let have a closed range. Then, and are orthogonally complemented. Lemma 4 ([
33], Theorem 2.2).
An operator is MP-invertible if and only if is closed. Lemma 5 ([
34], Theorem 1.3).
For any idempotents , let P and Q be two projections from to and , respectively. Then, and . Lemma 6. Let be an idempotent and such that . Then, .
Proof. As , for any , there exists such that . Thus, . □
3. Closedness and Orthogonal Complementability of the Sum of Adjointable Idempotents and Their Adjoints
For any idempotent
, let
P and
Q be two projections from
to
and
, respectively. Then, we have that
Based on Lemma 2, we deduce that
Therefore, to apply Halmos’ two projections theorem to the study of adjointable idempotents on Hilbert
-modules, according to (
1), it is necessary to investigate the orthogonal complementability of both
and
. Lemma 3 establishes that the closedness of
implies its orthogonal complementability. Consequently, we establish necessary and sufficient conditions for the closedness and orthogonal complementability of
in the following theorem.
Theorem 1. Let be an idempotent and P be a projection from to . Then, is closed if and only if is closed. Furthermore, is orthogonally complemented if and only if is orthogonally complemented.
Proof. and
can be derived from
Then,
This leads to
and it is clear that
is closed if and only if
is closed.
Now, suppose
is orthogonally complemented. Then, combined with (
12), we have the following:
Let
. Observe that
satisfies
, making it a projection. The conditions
further imply
. For any
such that
, then
, which means that
, yielding
. Consequently, based on (
13), we have that
is orthogonally complemented.
On the other hand, if
is orthogonally complemented, let
. Observe that
satisfies
, making it a projection. The conditions
further imply
. For any
such that
, then
. This means that
, yielding
. Consequently, based on (
13), we have that
is orthogonally complemented. □
Corollary 1. Let be an idempotent and Q be projection from to . Then, is closed if and only if is closed. Furthermore, is orthogonally complemented if and only if is orthogonally complemented.
Proof. Substituting , and P with and Q, respectively, in the proof of Theorem 1 immediately completes the argument. □
Remark 1. The necessary and sufficient conditions for closedness and orthogonal complementability established in Theorem 1 are significant. The following example demonstrates the existence of an idempotent operator and a projection P from to for which is not closed and is not orthogonally complemented.
Example 1. Let be the -algebra of all continuous complex-valued functions on [0, 1]. Let be the canonical Hilbert -module with -valued inner productwhere . Let be defined as , where and . Obviously, . Now, we prove that π is adjointable.Assume there exists such thatLet . Then,for any . Then, can be deduced by choosing , and can be deduced by choosing . Through direct compoutation, we can verify that Let P be a projection on . Then, it is clear that for any . Consequently, we havewhere Now, we prove that M is not closed. Let . Clearly,It follows that converges uniformly to , which means that neither M nor is closed. We now prove that is not orthogonally complemented.where . If there exists a projection Q from onto , then, for any , there is . Now, for , where for any , assume that . Then,As a is a continuous complex-valued function on [0,1], setting yields for any . This contradicts the condition , demonstrating that no orthogonal projection onto exists. Based on the preceding analysis, we formally define the class of idempotent operators for which Halmos’ two projections theorem is applicable.
Definition 1. denotes an idempotent such that both and are orthogonally complemented.
4. Moore–Penrose Inverse and Spectral Norm of the Sum of Adjointable Idempotents and Their Adjoints
For any , the Moore–Penrose inverse of T is denoted by . Lemma 4 establishes that the existence of the Moore–Penrose inverse for an idempotent operator is intimately connected to its closedness and the orthogonal complementability of its range. We now establish a sufficient condition for the orthogonal complementability of . Our analysis begins with operators , subsequently extending to the general case of arbitrary idempotents
Theorem 2. For any idempotent such that is closed, let P and Q be two projections from to and , respectively. Then, is closed, , and .
Proof. Clearly,
. As
, we obtain
which imply
. Let
and
(
) be defined as in (
2)–(
6). This gives
Lemma 5 implies
, which yields
. Combined with (
5), we have
Define
and the unitary operator
via Lemma 1. From
, it follows that
. Therefore,
is invertible when
. Note that
if
. We extend this by defining
when
. Set
Combining
(defined in (
7)) with the decompositions (
8), (
9), and Lemma 5, we obtain
Additionally,
Based on Lemma 4, the Moore–Penrose inverse of
exists. This yields the representation
It can be deduced, from
, that
Thus, equating the (2,1)- and (2,2)-entries, we obtain that
The Moore–Penrose inverse property
implies the matrix commutation relation
Performing matrix multiplication yields the following:
Equating the (1,2)-entries gives the following:
Equating the (1,1)- and (2,2)-entries gives
Furthermore, the Moore–Penrose inverse property
implies that
Left-multiplying by
and right-multiplying by
V at the (1,1)-entries of both matrices, we obtain
via (
19), implying
. Substituting
into (
20) yields
which gives
. Substituting
and
into the (1,2)-entries of the matrix yields the following:
Similarly, substituting
into (
19) produces the following:
Combining (
21), (
24), and (
25), we deduce that
b is the Moore–Penrose inverse of
.
Based on (
16) and (
17), we have
which establishes
Theorem 1 implies that
is closed.
Substituting (
23) into the (2,2)-position of (
22), yields the following:
Substituting (
21) and (
23) into the (1,1)-position of (
18) gives the following:
Combining the Moore–Penrose inverse property
with
, (
17), (
26), (
27), and (
28), we derive
□
Theorem 3. Let be an idempotent operator and P be the projection onto . Then, is MP-invertible if and only if is MP-invertible. The following identities hold: and Proof. We begin by establishing the necessary condition. If
is MP-invertible, inspired by Theorem 2, let
. We now prove that
A is the MP-inverse of
. Applying left-multiplication by
P and
to both sides of the Moore–Penrose inverse identity
we obtain
Then, combined with
and
, we have
and
Consequently,
Applying left-multiplication by
and right-multiplication by
to the Moore–Penrose identity
we obtain
Similarly, applying left-multiplication by
and right-multiplication by
P to the same identity
yields
Combined with (
31) and the Moore–Penrose identity
we have
which means
Together with (
29), this yields
Similarly, based on (
33) and (
32), we have the following:
Combining (
34) and (
30), we conclude that
is the Moore–Penrose inverse of
.
We now prove sufficiency for this theorem. For notational simplicity, we set
and, obviously,
. Based on the definition of the Moore–Penrose inverses, it is well-known that
Inspired by Theorem 2, let
We now prove that
B is the Moore–Penrose inverse of
. To prevent cumbersome expressions in subsequent derivations, let
. Combined with (
35)–(
38), we have
and
Right-multiplying both sides of (
39) by
yields the following:
Left-multiplying both sides of the above equation by
yields the following:
Then, left-multiplying both sides of the above equation by
, combined with (
35)–(
38), yields the following:
Based on (
39), (
40), and (
41), we have the following:
Furthermore, combined with (
35)–(
38), this yields the following:
Based on (
37) and (
42), we have the following:
As
, we obtain that
Combining (
43)–(
46), we establish that
B is the Moore–Penrose inverse of
. □
Corollary 2. Let be an idempotent operator and Q be the projection onto . Then, is MP-invertible if and only if is MP-invertible. The following identities hold:where Proof. Substituting , , and P with , , and Q, respectively, in Theorem 3 yields the proof. □
After investigating the Moore–Penrose inverse of , it is natural to consider its spectrum and norm. We denote the spectrum of by , which is used in the following theorem.
Theorem 4. For any idempotent operator , let P and Q denote the projections onto and , respectively. Then,and Proof. Based on Lemma 5, we obtain the following:
Combined with
, it follows that
Thus, we have
For any
, the operators
are both invertible, as
Define
Using
,
, and
, we compute
Left-multiplying
T by
and using (
47) yields the following:
Combining (
48), (
49), and (
50), we conclude that
is invertible if and only if
is invertible.
Let
be the unital commutative
-algebra generated by
I and
. Using the Gelfand transform ∧,
is isomorphic to
, with
Thus,
is invertible if and only if
which holds precisely when
Having excluded
and
in our derivation, we obtain
Finally, as
is self-adjoint and
, based on Lemma 5,
□
Remark 2. Note that 0 and 2 may not always be spectral points of . The necessary and sufficient conditions for 0 or 2 not being spectral points of are established below. Let P be the projection onto .
Case 1:
When , the operatoris invertible, asGiven , the operator satisfies the following conditions: Based on (
13)
, implies . Thus,Example: For , is it clear that . Case 2:
When , the operatoris invertible, with inversesGiven , the operator satisfies the following conditions: Thus,Example: For , it is clear that . Finally, for any where and either or , we can obtain 5. Decomposition of Adjointable Idempotent Operators on Hilbert -Modules
We perform a linear decomposition of the idempotent operators defined on (see Definition 1), then we provide representations for the resulting idempotent operators after decomposition. Finally, we extend these results to arbitrary linear idempotent operators (without requiring adjointability) on Hilbert -modules. In order to obtain the main results, we require that the set of all idempotents is a poset.
Definition 2 ([
35]).
The idempotents on a Hilbert -module form a poset if the order ≤ defined by if and only if holds for any idempotents on . Let . For any or , the following lemma provides a characterization of and , which yields a constructive method.
Lemma 7. For any , let P be a projection from to . Then,
- (i)
For any linear idempotents on , if and only if there exists linear idempotents on such that and .
- (ii)
For any linear idempotents on , if and only if there exists linear idempotents on such that and .
Proof. (i) Necessity: Let the subspaces
and projections
(
) be defined by (
2)–(
6). Let
Q be the projection onto
. Let
and the unitary operator
be defined by Lemma 1, with
V defined in (
14) and
in (
7). Based on (
15), we have
Represent
as
The condition
implies
Direct computation yields
subject to
The idempotence relation
gives
From (
15) and (
51), we can decompose
where
Based on (
54)–(
57),
is idempotent. The form in (
58) implies
, and, so,
. From (
15), we have
and, based on (
52) and (
53),
Thus,
.
Sufficiency: As
, we have
. With
,
As
and
, we conclude
.
(ii) Necessity: Set
The condition
implies the following:
This simplifies to the following:
The idempotence relation
yields the following:
We decompose
where
Based on (
61)–(
64),
is idempotent. The form in (
65) implies
, and, so,
. Using (
59),
and, so,
.
Sufficiency: As
, Lemma 6 gives the following:
As
and
, we conclude
. □
Motivated by the results obtained by Baksalary [
13], we pay attention to the decomposition of idempotents. We also use their results to prove a decomposition theorem. This is a parallel generalization of Baksalary’s results to Hilbert
-modules.
Lemma 8 (cf. [
13], Theorem 1).
Let be linear idempotents on such that . Let Ψ be their linear combination of the form with non-zero scalars and . Then, there are exactly four conditions:- (i)
;
- (ii)
;
- (iii)
;
- (iv)
.
Proof. The proof is the same as Baksalary’s proof in ([
13], Theorem 1). □
Now, for any linear idempotents , we try to characterize all idempotents on and non-zero scalars and such that , where . Obviously, , and must satisfy one condition in Lemma 8. The main theorem is as follows.
Theorem 5. For any , let P be a projection from to . Then,
- (i)
If , there exist two non-zero idempotents and on such that if and only if there exists non-zero idempotents on such that , , , and ;
- (ii)
If and , there exist two non-zero idempotents and on such that if and only if there exists non-zero idempotents on such that , , and ;
- (iii)
For , , and a non-zero scalar , the following are equivalent:
- (a)
There exist distinct non-zero idempotents on such that .
- (b)
There exist operators on satisfying
- i.
, ;
- ii.
, ;
- iii.
;
- iv.
with explicit expressions
Proof. (i) Assume there exist idempotents
on
such that
. Based on Lemma 8(i),
establishing
as
and
.
Conversely, for any idempotent
on
satisfying
and
,
which implies that
is idempotent. Thus,
with
.
Therefore, the decomposition exists if and only if there exists a non-zero idempotent on such that and . Using Lemma 7(ii), we determine that exists if and only if there exists a non-zero idempotent on such that , and . Note that 0 cannot be decomposed as , considering that only holds.
(ii) Suppose there exist idempotents
on
such that
. Based on Lemma 8(ii),
establishing
. As
, we have
.
Conversely, for any idempotents
on
satisfying
and
,
which implies that
is idempotent. Thus,
with
.
Therefore, the decomposition
exists if and only if there exists an idempotent operator
on
such that
Using Lemma 7(i), we obtain that the decomposition
exists if and only if there exists a non-zero idempotent
on
such that
and
. Note that
I cannot be decomposed as
, given that only
holds, and 0 cannot be decomposed as
, given that
.
(iii) Assume
are idempotents on
satisfying
. Set
. Based on Lemma 8(iv),
As
and
, we have
.
Conversely, for
on
with
,
, and
, define
for
. The condition implies
and, thus,
The decomposition is
.
Therefore, the decomposition exists if and only if there exists
on
satisfying
Note that here; otherwise, it is straightforward to derive a contradiction through computation.
To characterize
, let
Q be the projection onto
, with subspaces
and projections
(
) defined in (
2)–(
6). Let
be as in (
7), and
,
from Lemma 1, with
V defined in (
14). Represent
as
Using (
15),
From (
68), we obtain
Thus,
Define
and the operators
Then,
with
such that
and
. As
, we have
Furthermore,
implies
. Using (
59),
Thus,
.
Conversely, for on satisfying
, ;
, ;
;
;
set
. Then,
,
and, based on Lemma 6,
This completes the proof by (
69). □
Subsequently, we extend the above theorem to arbitrary idempotent operators (i.e., without requiring adjointability) on .
Theorem 6. For any Ψ on , let P be a projection from to . Then,
- (i)
If , there exist two non-zero idempotents and on such that if and only if there exists non-zero idempotents on such that , , , and ;
- (ii)
If and , there exist two non-zero idempotents and on such that if and only if there exists non-zero idempotents on such that , , and ;
- (iii)
For , , and a non-zero scalar , the following are equivalent:
- (a)
There exist distinct non-zero idempotents on such that .
- (b)
There exist operators on satisfying
- i.
, ;
- ii.
, ;
- iii.
;
- iv.
with explicit expressions
Proof. (i) Necessity: If there exists a decomposition
, then based on (
66), we have
. Now, define
. Then,
and
. To see that
, if
holds, then
, contradicting the hypothesis
. Moreover,
, as assuming equality would imply
and, consequently,
, contradicting the hypothesis
. Furthermore, based on calculation, we have
(i) Sufficiency: As
, we have
, which implies
. Consequently,
Moreover,
, if equality holds, then
would imply
, contradicting
. Similarly, to see
, if equality holds, then
would imply
, contradicting
.
(ii) Necessity: Assuming there exists a decomposition
, based on (
67), we have
, which implies
. Consequently,
Define
Clearly,
. Direct computation yields the following:
Moreover,
, as, if equality held, then
would imply
contradicting
. Finally, simple calculations give
(ii) Sufficiency: As
, we have
, which leads to
and
Thus,
Note that
as, if equality holds, then
would imply
, contradicting
.
(iii) Necessity: Assuming there exists a decomposition
, Lemma 8 implies
. Substituting
yields
From (
71), we directly deduce
Left-multiplying (
71) by
and
, respectively, gives
Define
These satisfy
Combining (
72) and (
73) yields the following
Note that
From this expression, it follows that
and, consequently,
(iii) Sufficiency: Let
. Then, based on
,
, and
, we have
and
Thus,
Note that
since
. Furthermore, to see
, if equality holds, then
would imply
, contradicting
. □
Remark 3. Let H be a Hilbert space. For any idempotent operator satisfying and , there always exists a decomposition of the forms described in Theorem 6(ii) and Theorem 6(iii). Consider the specific decomposition where and , giving .
Let P be the orthogonal projection from H onto . As and , we have and . Thus, there exist and with and . Let be the one-dimensional closed subspace generated by , and define by . Denote the composition by . Then, satisfies ,, , and . These properties, combined with Lemma 6, yieldNow, definewhere is any complex number not equal to 0 or 1. This gives the decompositionHowever, as demonstrated by the example in the next section, not every idempotent Π
admits a decomposition as the sum of two idempotents. Having completed our investigation into the linear decomposition of idempotent operators on Hilbert
-modules, we now examine their product decompositions. In [
24], Theorem 3.2, we established the factorization of an idempotent operator into a product of two idempotents on Hilbert spaces and derived explicit representations for such factorizations. We now extend these results to the framework of Hilbert
-modules.
Theorem 7. Let Ψ be an linear idempotent operator on and P be the projection onto . Then, for idempotents on if and only if there exist operators on and idempotents on satisfying the following:
- (1)
Range and null space conditions - (2)
- (3)
Factorization expressions
Proof. Necessity: Assuming there exists a product decomposition
, we have
. Based on Lemma 6, this implies
Using (
74) and Lemma 6, we derive the idempotency properties
The zero-product relations are established as follows:
The factorization expressions are established through the following derivations:
and
Sufficiency: The idempotency and product relations are verified through direct computation
□
6. Two Examples
To demonstrate the utility of Theorem 6 in decomposing idempotents, we provide the following example, which relies on a key lemma:
Lemma 9 ([
36], Theorem 8).
For projections on a finite-dimensional complex vector space M, the projection onto is , where denotes the Moore–Penrose inverse. Example 2. Consider the following idempotent matrix on :where and . Let P and Q be projections onto and , respectively, and let ) be defined as in (2)–(6), and and the unitary operator as defined in Lemma 1. Furthermore, let . Then, can be calculated using Lemma 9 combined with (59). In particular, we compute the following: Note that and .
Non-existence of additive decomposition: Assume there exists a decomposition as in Theorem 6(i). Then, there exist with . As is one-dimensional and spanned by , we have . Idempotence implies and, so, —a contradiction.
MATLAB R2022a(9.12) verification confirmed that no solution exists for , , .
Existence of alternative decompositions: As , Theorem 6(ii) gives the following: Considering Theorem 6(iii), let and This yields .
Throughout
Section 3 and
Section 4, we focused exclusively on
adjointable idempotent operators on Hilbert
-modules. This restriction is necessary due to a fundamental difference between Hilbert
-modules and Hilbert spaces: unlike operators on Hilbert spaces, idempotent operators on Hilbert
-modules may not admit an adjoint. A simple counterexample illustrating this phenomenon is presented below.
Example 3. Let be the algebra of complex-valued continuous functions on . Let be a closed ideal in B. Define the Hilbert B-module with the following inner product: Define the idempotent operator as follows:Verification of idempotencyVerification thatAhas no adjoint Assume there exists an adjoint satisfying for all . The left side is For the right side, let . Then, Equating both sides yields Taking gives For , we have . By continuity, . Substituting into (76) yields the following: Thus, , implying that . Note that requires , as requires the first component in I. However, need not belong to I; for example, the constant function satisfies . Hence, does not map into , giving a contradiction.