1 Introduction and preliminaries

The class of absolutely norm attaining operators (or shortly, \({\mathcal {AN}}\)-operators) between complex Hilbert spaces was introduced and several important class of examples and properties of these operators were discussed by Carvajal and Neves in [3]. Later, a structure of these operators on separable Hilbert spaces was proposed in [11]. But, an example of \({\mathcal {AN}}\)-operator which does not fit into the characterization of [11] was given in [10] and the authors discussed the structure of positive \({\mathcal {AN}}\)-operators between arbitrary Hilbert spaces. In this article, first, we give necessary and sufficient conditions for an operator to be positive and \({\mathcal {AN}}\). In fact, we show that a bounded operator T defined on an infinite dimensional Hilbert space is positive and \({\mathcal {AN}}\) if and only if there exists a unique triple \((K,F,\alpha )\), where K is a positive compact operator, F is a positive finite rank operator, \(\alpha \) is a positive real number such that \(T=K-F+\alpha I\) and \(KF=FK=0,\; F\le \alpha I\) (see Theorem 2.5). In fact, here \(\alpha =m_e(T)\), the essential minimum modulus of T. This is an improvement of [10, Theorem 5.1]. Using this result, we give explicit structure of self-adjoint and \({\mathcal {AN}}\)-operators as well as normal and \({\mathcal {AN}}\)-operators. Finally, we also obtain structure of general \({\mathcal {AN}}\)-operators. In the process, we also prove several important properties of \({\mathcal {AN}}\)-operators. All these results are new.

We organize the article as follows: In the remaining part of this section, we explain the basic terminology, notations and necessary results that will be needed for proving main theorems. In section 2, we give a characterization of positive \(\mathcal { AN}\)-operators and prove several important properties. In section 3, we discuss the structure of self-adjoint and normal \({\mathcal {AN}}\)-operators and in section 4, we discuss about the general \({\mathcal {AN}}\)-operators.

Throughout the article we consider complex Hilbert spaces which will be denoted by \(H,H_1,H_2\), etc. The inner product and the induced norm are denoted by \(\langle , \rangle \) and \(\Vert \cdot \Vert \) respectively. The unit sphere of a closed subspace M of H is denoted by \(S_M:={\{x\in M: \Vert x\Vert =1}\}\) and \(P_M\) denotes the orthogonal projection \(P_M:H\rightarrow H\) with range M. The identity operator on M is denoted by \(I_M\).

A linear operator \(T: H_1\rightarrow H_2\) is said to be bounded if there exists a number \(k>0\) such that \(\Vert Tx\Vert \le k \Vert x\Vert \) for all \(x\in H_1\). If T is bounded, the quantity \(\Vert T\Vert =\sup {\{\Vert Tx\Vert : x\in S_{H_1}}\}\) is finite and is called the norm of T. We denote the space of all bounded linear operators between \(H_1\) and \(H_2\) by \({\mathcal {B}}(H_1,H_2)\). In case if \(H_1=H_2=H\), then \(\mathcal B(H_1,H_2)\) is denoted by \({\mathcal {B}}(H)\). For \(T\in \mathcal B(H_1,H_2)\), there exists a unique operator denoted by \(T^*:H_2\rightarrow H_1\) satisfying

$$\begin{aligned} \langle Tx,y\rangle =\langle x, T^*y\rangle \; \text {for all}\; x\in H_1 \; \text {and} \; \text {for all}\; y\in H_2. \end{aligned}$$

This operator \(T^*\) is called the adjoint of T. The null space and the range spaces of T are denoted by N(T) and R(T) respectively.

Let \(T\in {\mathcal {B}}(H)\). Then T is said to be normal if \(T^*T=TT^*\), self-adjoint if \(T=T^*\). If \(\langle Tx,x\rangle \ge 0\) for all \(x\in H\), then T is called positive. It is well known that for a positive operator T, there exists a unique positive operator \(S\in {\mathcal {B}}(H)\) such that \(S^2=T\). We write \(S=T^{\frac{1}{2}}\) and it is called as the positive square root of T.

If \(S,T\in {\mathcal {B}}(H)\) are self-adjoint and \(S-T\ge 0\), then we write this by \(S\ge T\).

If \(P\in {\mathcal {B}}(H)\) is such that \(P^2=P\), then P is called a projection. If N(P) and R(P) are orthogonal to each other, then P is called an orthogonal projection. It is a well known fact that a projection P is orthogonal if and only if it is self-adjoint if and only if it is normal.

We call an operator \(V\in {\mathcal {B}}(H_1,H_2)\) to be an isometry if \(\Vert Vx\Vert =\Vert x\Vert \) for each \(x\in H_1\). An operator \(V\in {\mathcal {B}}(H_1,H_2)\) is said to be a partial isometry if \(V|_{N(V)^{\bot }}\) is an isometry. That is \(\Vert Vx\Vert =\Vert x\Vert \) for all \(x\in N(V)^{\bot }\). If \(V\in {\mathcal {B}}(H)\) is isometry and onto, then V is said to be a unitary operator.

In general, if \(T\in {\mathcal {B}}(H_1,H_2)\), then \(T^*T\in \mathcal B(H_1)\) is positive and \(|T|:=(T^*T)^{\frac{1}{2}}\) is called the modulus of T. In fact, there exists a unique partial isometry \(V\in {\mathcal {B}}(H_1,H_2)\) such that \(T=V|T|\) and \(N(V)=N(T)\). This factorization is called the polar decomposition of T.

If \(T\in {\mathcal {B}}(H)\), then \(T=\frac{T+T^*}{2}+i(\frac{T-T^*}{2i})\). The operators \(\mathrm{Re}(T):=\frac{T+T^*}{2}\) and \(\mathrm{Im}(T):=\frac{T-T^*}{2i}\) are self-adjoint and are called the real and the imaginary parts of T respectively.

A closed subspace M of H is said to be invariant under \(T\in {\mathcal {B}}(H)\) if \(TM\subseteq M\) and reducing if both M and \(M^\bot \) are invariant under T.

For \(T\in {\mathcal {B}}(H)\), the set

$$\begin{aligned} \rho (T):={\{\lambda \in \mathbb C: T-\lambda I \text { is invertible and} \;(T-\lambda I)^{-1}\in {\mathcal {B}}(H) }\} \end{aligned}$$

is called the resolvent set and the complement \(\sigma (T)={\mathbb {C}}\setminus \rho (T)\) is called the spectrum of T. It is well known that \(\sigma (T)\) is a non empty compact subset of \({\mathbb {C}}\). The point spectrum of T is defined by

$$\begin{aligned} \sigma _p(T)={\{\lambda \in {\mathbb {C}}: T-\lambda I\ \text {is not one-to-one}}\}. \end{aligned}$$

Note that \(\sigma _{p}(T)\subseteq \sigma (T)\).

A self-adjoint operator \(T\in {\mathcal {B}}(H)\) is positive if and only if \(\sigma (T)\subseteq [0,\infty )\).

A bounded linear operator \(T:H_1\rightarrow H_2\) is called finite rank if R(T) is finite dimensional. The space of all finite rank operators between \(H_1\) and \(H_2\) is denoted by \({\mathcal {F}}(H_1,H_2)\) and we write \({\mathcal {F}}(H,H)={\mathcal {F}}(H)\). If \(T\in {\mathcal {B}}(H_1,H_2)\), then T is said to be compact if for every bounded set S of \(H_1\), the set T(S) is pre-compact in \(H_2\). Equivalently, for every bounded sequence \((x_n)\) of \(H_1\), \((Tx_n)\) has a convergent subsequence in \(H_2\). We denote the set of all compact operators between \(H_1\) and \(H_2\) by \({\mathcal {K}}(H_1,H_2)\). In case if \(H_1=H_2=H\), then \({\mathcal {K}}(H_1,H_2)\) is denoted by \({\mathcal {K}}(H)\).

All the above mentioned basics of operator theory can be found in [4, 8, 12, 13].

An operator \(T\in {\mathcal {B}}(H_1,H_2)\) is said to be norm attaining if there exist \(x\in S_{H_1}\) such that \(\Vert Tx\Vert =\Vert T\Vert \). We denote the class of norm attaining operators by \(\mathcal N(H_1,H_2)\). It is known that \({\mathcal {N}}(H_1,H_2)\) is dense in \({\mathcal {B}}(H_1,H_2)\) with respect to the operator norm of \(\mathcal B(H_1,H_2)\). We refer [5] for more details on this topic.

We say \(T\in {\mathcal {B}}(H_1,H_2)\) to be absolutely norm attaining or \({{\mathcal {AN}}}\)-operator (shortly), if \(T|_M\), the restriction of T to M, is norm attaining for every non zero closed subspace M of \(H_1\). That is, \(T|_M\in {\mathcal {N}}(M, H_2)\) for every non zero closed subspace M of \(H_1\) [3]. This class contains \({\mathcal {K}}(H_1,H_2)\), and the class of partial isometries with finite dimensional null space or finite dimensional range space.

We have the following characterization of norm attaining operators.

PROPOSITION 1.1

[3, Proposition 2.4]

Let \(T\in {\mathcal {B}}(H)\) be self-adjoint. Then

  1. (1)

    \(T\in {\mathcal {N}}(H)\) if and only if either \(\Vert T\Vert \in \sigma _p(T)\) or \(-\Vert T\Vert \in \sigma _p(T),\)

  2. (2)

    if \(T\ge 0\), then \(T\in {\mathcal {N}}(H)\) if and only if \(\Vert T\Vert \in \sigma _p(T)\).

For \(T\in {{\mathcal {B}}}(H_1,H_2)\), the quantity

$$\begin{aligned} m(T):=\inf {\{\Vert Tx\Vert : x\in S_{H_1}}\} \end{aligned}$$

is called the minimum modulus of T. If \(H_1=H_2=H\) and \(T^{-1}\in {\mathcal {B}}(H)\), then \(m(T)=\dfrac{1}{\Vert T^{-1}\Vert }\) (see [1, Theorem 1] for details).

The following definition is available in [9] for densely defined closed operators (not necessarily bounded) on a Hilbert space, and this holds true automatically for bounded operators.

DEFINITION 1.2

[9, Definition 8.3 p. 178]

Let \(T=T^*\in \mathcal B(H)\). Then the discrete spectrum \(\sigma _d(T)\) of T is defined as the set of all eigenvalues of T with finite multiplicities which are isolated points of the spectrum \(\sigma (T)\) of T. The complement set \(\sigma _{\mathrm{ess}}(T)=\sigma (T)\setminus \sigma _d(T)\) is called the essential spectrum of T.

By the Weyl’s theorem we can assert that if \(T=T^*\) and \(K=K^*\in {\mathcal {K}}(H)\), then \(\sigma _{\mathrm{ess}}(T+K)=\sigma _{\mathrm{ess}}(T)\) (see [9, Corollary 8.16, p. 182] for details). If H is a separable Hilbert space, the essential minimum modulus of T is defined to be \(m_e(T):=\inf {\{\lambda : \lambda \in \sigma _{\mathrm{ess}}(|T|)}\}\) (see [1] for details). The same result in the general case is dealt in [2].

Let \(H=H_1\oplus H_2\) and \(T\in {\mathcal {B}}(H)\). Let \(P_j:H\rightarrow H\) be an orthogonal projection onto \(H_j\) for \(j=1,2\). Then \( T=\left( \begin{array}{cc} T_{11} &{}T_{12} \\ T_{21} &{}T_{22} \\ \end{array} \right) \), where \(T_{ij}:H_j\rightarrow H_i\) is the operator given by \(T_{ij}=P_iTP_j|_{H_j}\). In particular, \(T(H_1)\subseteq H_1\) if and only if \(T_{12}=0\). Also, \(H_1\) reduces T if and only if \(T_{12}=0=T_{21}\) (for details, see [4, 13]).

2 Positive \({\mathcal {AN}}\)-operators

In this section, we describe the structure of operators which are positive and satisfy the \({\mathcal {AN}}\)-property. First, we recall the results that are necessary for proving our results.

Theorem 2.1

[10, Theorem 5.1]. Let H be a complex Hilbert space of arbitrary dimension and let P be a positive operator on H. Then P is an \({\mathcal {AN}}\)-operator iff P is of the form \(P = \alpha I + K + F\), where \(\alpha \ge 0\), K is a positive compact operator and F is a self-adjoint finite rank operator.

Theorem 2.2

[10, Theorem 3.8]. Let \(T\in B(H)\) be positive and \(T\in {\mathcal {AN}}(H)\). Then

$$\begin{aligned} T=\displaystyle \sum _{\alpha \in \Lambda } \beta _{\alpha }v_{\alpha }\otimes v_{\alpha }, \end{aligned}$$
(2.1)

where \({\{v_{\alpha }:\alpha \in \Lambda }\}\) is an orthonormal basis consisting of entirely eigenvectors of T and for every \(\alpha \in \Lambda \), \(Tv_{\alpha }=\beta _{\alpha }v_{\alpha }\) with \(\beta _{\alpha }\ge 0\) such that

  1. (1)

    for every non empty set \(\Gamma \) of \(\Lambda \), we have

    $$\begin{aligned} \sup {\{\beta _{\alpha }:\alpha \in \Gamma }\}=\max {\{\beta _{\alpha }:\alpha \in \Gamma }\}\mathrm{,} \end{aligned}$$
  2. (2)

    the spectrum \(\sigma (T)=\overline{{\{\beta _{\alpha }:\alpha \in \Lambda }\}}\) has at most one limit point. Moreover, this unique limit point (if exists) can only be the limit of an increasing sequence in the spectrum,

  3. (3)

    the set \({\{\beta _{\alpha }:\alpha \in \Lambda }\}\) of eigenvalues of T, without counting multiplicities, is countable and has atmost one eigenvalue with infinite multiplicity,

  4. (4)

    if \(\sigma (T)\) has both, i.e., a limit point and an eigenvalue with infinite multiplicity, then they must be the same.

Here \((v_{\alpha }\otimes v_{\alpha })(x)=\langle x,v_{\alpha }\rangle v_{\alpha }\) for each \(\alpha \in \Lambda \) and for each \(x\in H\).

Remark 2.3

In Theorem 2.1, the standing hypothesis is that the operator T is positive. In particular, if K is a positive compact operator, F is a self-adjoint finite rank operator and \(\alpha \ge 0\), it is interesting to know when \(T:=K+F+\alpha I\) is positive? We will answer this question later in this article and also discuss about the uniqueness of the representation of T given in Theorem 2.1.

Lemma 2.4

Let \(S,T\in {\mathcal {B}}(H)\) be positive such that \(S\le T\). Then \(N(T)\subseteq N(S)\).

Proof

If \(x\in H\), then \(\Vert S^{\frac{1}{2}}x\Vert ^2=\langle Sx,x\rangle \le \langle Tx,x\rangle =\Vert T^{\frac{1}{2}}x\Vert ^2\). By observing the fact that for any \(A\in {\mathcal {B}}(H)\) with \(A\ge 0\), \(N(A^{\frac{1}{2}})=N(A)\), the conclusion follows. \(\square \)

Theorem 2.5

Let H be an infinite dimensional Hilbert space and \(T\in \mathcal B(H)\). Then the following statements are equivalent:

  1. (1)

    \(T\in {{\mathcal {AN}}}(H)\) and positive,

  2. (2)

    there exists a unique triple \((K,F,\alpha )\), where

    1. (a)

      \(K\in {\mathcal {K}}(H)\) is positive, \(\alpha \ge 0\),

    2. (b)

      \(F\in {\mathcal {F}}(H)\) and \(0\le F\le \alpha I\),

    3. (c)

      \(KF=0\),

    such that \(T=K-F+\alpha I\).

Proof

Proof of (1) \(\Rightarrow \) (2). By Theorem 2.1, \(T=K'-F'+\alpha I\), where \(K'\in \mathcal K(H)\) is positive, \(F'=F'^{*}\in {\mathcal {F}}(H)\) and \(\alpha \ge 0\). As \(K'-F'\) is compact, self-adjoint, there exists an orthonormal set \({\{\phi _n:n\in {\mathbb {N}}}\}\) of eigenvectors corresponding to the eigenvalues \({\{\lambda _n:n\in {\mathbb {N}}}\}\). That is, \((K'-F')(\phi _n)=\lambda _n\phi _n\) for each \(n\in {\mathbb {N}}\). By [10, Lemma 4.8], \(K'-F'\) can have at most finitely many negative eigenvalues. Without loss of generality, assume that \(\lambda _1,\lambda _2,\ldots , \lambda _k\) are those negative eigenvalues. Define \(F(x)= \sum _{n=1}^{k}(-\lambda _n)\langle x,\phi _{n}\rangle \phi _n\) and \(K(x)= \sum _{n=k+1}^{\infty }\lambda _n\langle x,\phi _{n}\rangle \phi _n\) for all \(x\in H\). Then clearly, \(FK=KF=0\) and F is a positive finite rank operator and K is a positive compact operator. Also, \(K'-F'=K-F\) and hence \(T=\alpha I+K-F\).

Next, \(TF=(\alpha I-F)F=FT\). Since T and F are positive, it follows that TF is positive. Let \(\lambda \in \sigma (F)\). Then \(\lambda \ge 0\) and since \(FT\ge 0\), by the spectral mapping theorem, we have that \(\lambda (\alpha -\lambda )\ge 0\). From this, we can conclude that \(\alpha -\lambda \ge 0\) for each \(\lambda \in \sigma (F)\). As \(\alpha I-F\) is self-adjoint and \(\sigma (\alpha I-F)\subseteq [0,\infty )\), \(\alpha I-F\) must be positive. This concludes that \(F\le \Vert F\Vert I\le \alpha I\).

Next, we show that the triple satisfying the given conditions is unique. Suppose there exists two triples \((K_1,F_1,\alpha _1), (K_2,F_2,\alpha _2 )\) satisfying the stated conditions. We prove this by considering all possible cases.

Case 1. \(\alpha _1=0\). In this case, \(F_1=0\). Hence \(K_1=T=K_2-F_2+\alpha _2I\). This shows that \(\alpha _2I=K_1-K_2+F_2\), a compact operator. Since H is infinite dimensional, it follows that \(\alpha _2=0\). Thus \(F_2=0\). Hence we can conclude that \(K_1=K_2\).

Case 2. \(F_1=0\), \(\alpha _1>0\). In this case,

$$\begin{aligned} K_1+\alpha _1I=K_2-F_2+\alpha _2I. \end{aligned}$$
(2.2)

Then \((\alpha _2-\alpha _1)I=(K_1-K_2)+F_2\) , a compact operator. Hence \(\alpha _1=\alpha _2\).

Now equation (2.2) can be written as \(K_2=F_2+K_1\ge F_2\). By Lemma  2.4, we have that \(N(K_2)\subseteq N(F_2)\). But, by the condition \(K_2F_2=0\), we have \(R(F_2)\subseteq N(K_2)\), and hence \(R(F_2)\subseteq N(F_2)\). Thus, \(F_2=0\). From this, we can conclude that \(K_1=K_2\).

Case 3. \( K_1=0\), \(F_1\ne 0\), \(\alpha _1>0\). We have \(F_1+\alpha _1 I=K_2-F_2+\alpha _2 I\). Using the same argument as in the above cases, we can conclude that \(\alpha _1=\alpha _2\). Thus we have \(F_2=K_2+F_1\ge K_2\). Now, by Lemma 2.4, \(N(F_2)\subseteq N(K_2)\). But by the property \(K_2F_2=0\), it follows that \(R(F_2)\subseteq N(K_2)\). Hence \(H=N(F_2)\oplus R(F_2)\subseteq N(K_2)\). This shows that \(K_2=0\). Finally, using this we can get \(F_1=F_2\).

Case 4. \(K_1\ne 0\), \(F_1\ne 0\), \(\alpha _1>0\). We can prove \(\alpha _1=\alpha _2\) by arguing as in the earlier cases. With this, we have

$$\begin{aligned} K_1-F_1=K_2-F_2. \end{aligned}$$
(2.3)

As \(F_1\) commute with \(K_1\) and \(F_1\), it commute with \(K_2-F_2\). So \(F_1\) must commute with \((K_2-F_2)^2=K_2^2+F_2^2=(K_2+F_2)^2\). Thus, it commute with \(K_2+F_2\). Hence we can conclude that \(F_1\) commute with both \(K_2\) and \(F_2\). Since \(N(F_1)\) is invariant under \(K_1\) and \(F_1\), by equation (2.3), \(N(F_1)\) is invariant under \(K_2-F_2\).

Now if \(x\in N(F_1)\), then by equation (2.3), we have \((K_2-K_1)x=F_2x\). Using the fact that \(F_2\ge 0\), we can conclude that \(K_2\ge K_1\) on \(N(F_1)\). We also show that this will happen on \(R(F_1)\).

For \(x\in H\), we have \(F_1x\in R(F_1)\). Now,

$$\begin{aligned} \langle (F_2-F_1)(F_1x),F_1x\rangle =\langle (K_2-K_1)(F_1x),F_1x\rangle =\langle K_2(F_1x),F_1x\rangle \ge 0. \end{aligned}$$

This shows that \(K_2-K_1=F_2-F_1\ge 0\) on \(R(F_1)\). Combining with the earlier argument, we can conclude that \(K_1\le K_2\). Now, interchanging the roles of \(K_1\) and \(K_2\), we can conclude that \(K_2\le K_1\) and hence \(K_1=K_2\). By equation (2.3), we can conclude that \(F_1=F_2\).

Proof of (2) \(\Rightarrow \) (1). If \(T=K-F+\alpha I\), where \(K\in {\mathcal {K}}(H)\) is positive, \(F\in {\mathcal {F}}(H)\) is positive, \(\alpha \ge 0\) and \(KF=0\). Then by Theorem 2.1, \(T\in {{\mathcal {AN}}}(H)\). Since \(K\ge 0\) and \(-F+\alpha I\ge 0\), T must be positive. \(\square \)

Remark 2.6

Let T be as in Theorem 2.5. Then we have the following:

  1. (1)

    if \(\alpha =0\), then \(F=0\) and hence \(T=K\). In this case \(\sigma _{\mathrm{ess}}(T)={\{\alpha }\}\),

  2. (2)

    if \(\alpha >0\) and \(F=0\), then \(T=K+\alpha I\). In this case, \(\sigma _{\mathrm{ess}}(T)={\{\alpha }\}\) and \(m_{e}(T)=\alpha =m(T)\),

  3. (3)

    if \(\alpha >0,\; K=0\) and \(F\ne 0\), then \(T=\alpha I-F\). In this case also, \(\sigma _{\mathrm{ess}}(T)={\{\alpha }\}\) and \(m_{e}(T)=\alpha \),

  4. (4)

    if \(\alpha >0,\; F\ne 0\) and \(K\ne 0\), then by the Weyl’s theorem, \(\sigma _{\mathrm{ess}}(T)={\{\alpha }\}\) and \(m_{e}(T)=\alpha \),

  5. (5)

    if \(\alpha =0\) and \(K=0\), then \(T=0\),

  6. (6)

    if N(T) is infinite dimensional, then 0 is an eigenvalue with infinite multiplicity and hence \(\alpha =0\), by Theorem 2.2. In this case, \(F=0\) and hence \(T=K\).

Remark 2.7

If we take \(F=0\) in Theorem 2.5, then we get the structure obtained in [11].

Here we prove some important properties of \({\mathcal {AN}}\)-operators.

PROPOSITION 2.8

Let \(T=K-F+\alpha I\), where \(K\in {\mathcal {K}}(H)\) is positive, \(F\in {\mathcal {F}}(H)\) is positive with \(KF=0\) and \(F\le \alpha I\). If \(\alpha >0\), then the following statements hold:

  1. (1)

    R(T) is closed,

  2. (2)

    N(T) is finite dimensional,

  3. (3)

    \(N(T)\subseteq N(K)\),

  4. (4)

    \(Fx=\alpha x\) for all \(x\in N(T)\). Hence \(N(T)\subseteq R(F)\). In this case, \(\Vert F\Vert =\alpha \),

  5. (5)

    T is one-to-one if and only if \(\Vert F\Vert <\alpha \),

  6. (6)

    T is Fredholm and \(m_e(T)=\alpha \).

Proof

Proof of (1). Since \(K-F\) is a compact operator, R(T) is closed. Here we have used the fact that for any \(A\in {\mathcal {K}}(H)\) and \(\lambda \in {\mathbb {C}}\setminus {\{0}\}\), \(R(A+\lambda I)\) is closed.

Proof of (2). Let \(x\in N(T)\). Then

$$\begin{aligned} (K-F)x=-\alpha x, \end{aligned}$$
(2.4)

that is, \(\alpha I_{N(T)}\) is compact. This concludes that N(T) is finite dimensional.

Proof of (3). Let \(x\in N(T)\). Multiplying Equation (2.4) by K and using the fact that \(KF=FK=0\), we have \(K^2x=-\alpha Kx\). If \(Kx\ne 0\), then \(-\alpha \in \sigma _p(K)\) contradicts the positivity of K. Hence \(Kx=0\).

Proof of (4). Clearly, if \(Tx=0\), then by (3), we have \(Fx=\alpha x\). This also concludes that \(N(T)\subseteq R(F)\) and \(\Vert F\Vert =\alpha \).

Proof of (5). If T is not one-to-one, then \(Fx=\alpha x\) for \(x\in N(T)\) by (3). Suppose T is one-to-one and \(\Vert F\Vert =\alpha \). Since F is norm attaining by Proposition 1.1, there exists \(x\in S_H\) such that \(Fx=\alpha x\). Then \(Tx=Kx-Fx+\alpha x=Kx\). But \(KF=0\) implies that \(x\in N(K)\). So, \(Tx=Kx\)=0. By the injectivity of T, we have that \(x=0\). This contradicts the fact that \(x\in S_H\). Hence \(\Vert F\Vert <\alpha \).

Proof of (6). Note that \(\sigma _\mathrm{ess}(T)={\{\alpha }\}\) by the Weyl’s theorem on essential spectrum. Hence \(m_e(T)=\alpha =m_e(T^*)\). Now T is Fredholm operator by [1, Theorem 2] with index zero. \(\square \)

Theorem 2.9

Let \(T\in {\mathcal {B}}(H)\) and positive. Then \(T\in {{\mathcal {AN}}}(H)\) if and only if \(T^2\in {{\mathcal {AN}}}(H)\).

Proof

First we assume that \(T\in {{\mathcal {AN}}}(H).\) Then there exists a triple \((K, F,\alpha )\) as in Theorem 2.5(2). Then \(T^2=K_1-F_1+\beta I\), where \(K_1=K^2+2\alpha K\) is a positive compact operator, \(F_1=2\alpha F-F^2=(2\alpha I-F)F\) and \(\beta =\alpha ^2\). Clearly, \(F_1\ge 0\) as it is the product of two commuting positive operators. Also, \(F_1\in {\mathcal {F}}(H)\). Next, we show that \(F_1\le \alpha ^2I\). Clearly, \(\alpha ^2I-F_1\) is self-adjoint and \(\alpha ^2I-F_1=(\alpha I-F)^2\ge 0\). It can be easily verified that \(K_1F_1=0\). So, \(T^2\) is also in the same form. Hence by Theorem 2.5, \(T^2\in {{\mathcal {AN}}}(H)\).

Now, let \(T^2\in {{\mathcal {AN}}}(H)\). Then by Theorem 2.5, \(T^2= K-F + \alpha I\), where \(K\in {{\mathcal {K}}}(H)\) is positive, \(F\in {\mathcal {F}}(H)\) is positive with \(FK=KF=0\) and \(F\le \alpha I \). If \(\alpha > 0,\) then \((T-\sqrt{\alpha } I)(T+\sqrt{\alpha } I)= K-F\). Since T is positive \(T+\sqrt{\alpha } I\) is a positive invertible operator. Hence \(T-\sqrt{\alpha } I= (K-F)(T+\sqrt{\alpha } I)^{-1}\). Hence there is a positive compact operator, namely \(K_1 = K (T+\sqrt{\alpha })^{-1}\) and a finite rank positive operator, namely \(F_1 = F (T+\sqrt{\alpha } I)^{-1},\) such that \(T-\sqrt{\alpha } I= K_1 - F_1.\) Hence \(T= K_1-F_1 + \sqrt{\alpha }I\). Also, note that since F and K commute with \(T^2\), hence commutes with T. Thus, we can conclude that \(F_1K_1=0\). Finally,

$$\begin{aligned} \Vert F_1\Vert \le \Vert F\Vert \, \Vert (T+\sqrt{\alpha } I)^{-1}\Vert&\le \alpha \, \frac{1}{m(T+\sqrt{\alpha } I)}\\&=\frac{\alpha }{\sqrt{\alpha }+m(T)} \\&\le \frac{\alpha }{\sqrt{\alpha }}=\sqrt{\alpha }. \end{aligned}$$

In the third step of the above inequalities, we used the fact that \(m(T+\sqrt{\alpha } I)=\sqrt{\alpha }+m(T)\), which follows by [11, Proposition 2.1].

If \(\alpha =0\), then clearly \(F=0\) and hence \(T^2=K\). So, \(T=K^{\frac{1}{2}}\), a compact operator which is clearly an \({\mathcal {AN}}\)-operator. \(\square \)

COROLLARY 2.10

Let \(T\in {\mathcal {B}}(H)\) and positive. Then \(T\in {{\mathcal {AN}}}(H)\) if and only if \(T^{\frac{1}{2}}\in {{\mathcal {AN}}}(H)\).

Proof

Let \(S=T^{\frac{1}{2}}\). Then \(S\ge 0\). The conclusion follows by Theorem 2.9. \(\square \)

COROLLARY 2.11

Let \(T\in {\mathcal {B}}(H_1,H_2)\). Then \(T\in {{\mathcal {AN}}}(H_1,H_2)\) if and only \(T^*T\in {{\mathcal {AN}}}(H_1)\).

Proof

Proof follows from the following arguments: \(T^*T\in {{\mathcal {AN}}}(H_1)\Leftrightarrow |T|^2\in {{\mathcal {AN}}}(H_1)\Leftrightarrow |T|\in {{\mathcal {AN}}}(H_1) \Leftrightarrow T \in {{\mathcal {AN}}}(H_1, H_2).\) \(\square \)

We have the following consequence.

Theorem 2.12

Let \(T\in {{\mathcal {AN}}}(H)\) be self-adjoint and \(\lambda \) be a purely imaginary number. Then \(T \pm \lambda I \in {\mathcal {AN}}(H)\).

Proof

Let \(S=T \pm \lambda I\). Then \(S^*S=T^2+|\lambda |^2I =K-F+(\alpha +|\lambda |^2)I\), where the triple \((K,F,\alpha )\) satisfy conditions Theorem 2.5(2) applied to \(T^2\). Hence by Corollary 2.11, \(S\in {\mathcal {AN}}(H)\). \(\square \)

The following result is well known.

Lemma 2.13

Let \(S,T\in {\mathcal {B}}(H)\) be such that \(S^{-1},T^{-1}\in \mathcal B(H)\). Then \(S^{-1}-T^{-1}=T^{-1}(T-S)S^{-1}\).

Theorem 2.14

Let \(T=K-F+\alpha I\), where \((K, F, \alpha )\) satisfy conditions of Theorem 2.5(2). Then

  1. (1)

    R(F) reduces T,

  2. (2)

    \(T=\left( \begin{array}{cc} K_0+\alpha I_{N(F)}&{} 0 \\ 0 &{} \alpha I_{R(F)}-F_0 \\ \end{array} \right) \), where \(K_0=K|_{N(F)}\) and \(F_0=F|_{R(F)}\),

  3. (3)

    if T is one-to-one and \(\alpha >0\), then \(T^{-1}\in {\mathcal {B}}(H)\) and

    $$\begin{aligned} T^{-1}=\left( \begin{array}{cc} \alpha ^{-1}I_{N(F)}-\alpha ^{-1}K_0(K_0+\alpha I_{N(F)})^{-1} &{}0 \\ 0 &{} \alpha ^{-1}I_{R(F)}+\alpha ^{-1}F_0(\alpha I_{R(F)}-F_0)^{-1} \\ \end{array} \right) . \end{aligned}$$

Proof

Proof of (1). First note that \(T\ge 0\) and \(T\in {\mathcal {AN}}(H)\). Let \(y=Fx\) for some \(x\in H\). Then \(Ty=TFx=(K-F+\alpha I)Fx=(\alpha I-F)(Fx)=F(\alpha I-F)x\in R(F)\). This shows that R(F) is invariant under T. As T is positive, it follows that R(F) is a reducing subspace for T.

Proof of (2). First, we show that \(K_0\) is a map on N(F). For this we show that N(F) is invariant under K. If \(x\in N(F)\), then \(FKx=0\) since \(FK=0\). This proves that N(F) is invariant under K. Thus \(K_0\in {\mathcal {K}}(N(F))\). Also, clearly, R(F) is invariant under F. Thus \(F_0: R(F)\rightarrow R(F)\) is a finite dimensional operator. With respect to the pair of subspaces (N(F), R(F)), K has the decomposition

$$\begin{aligned} \left( \begin{array}{ll} K_0 &{} \quad 0 \\ 0 &{}\quad 0 \\ \end{array} \right) . \end{aligned}$$

Similarly the operators F and \(\alpha I\) has the following block matrix forms:

$$\begin{aligned} \left( \begin{array}{ll} 0 &{}\quad 0 \\ 0 &{}\quad F_0 \\ \end{array} \right) \; \; \text {and}\;\; \left( \begin{array}{ll} \alpha I_{N(F)} &{}\quad 0 \\ 0 &{}\quad \alpha I_{R(F)} \\ \end{array} \right) . \end{aligned}$$

With these representations of KF and \(\alpha I\), by definition, T can be represented as in (2).

Proof of (3). By Proposition 2.8(1), R(T) is closed. As T is one-to-one, T is bounded below. Since T is positive, \(T^{-1}\in {\mathcal {B}}(H)\). In this case, \(\Vert F_0\Vert =\Vert F\Vert <\alpha \), by (5) of Proposition 2.8. Hence we have

$$\begin{aligned} T^{-1}=\left( \begin{array}{cc} (K_0+\alpha I_{N(F)})^{-1} &{} 0 \\ 0 &{} (\alpha I_{R(F)}-F_0)^{-1} \\ \end{array} \right) . \end{aligned}$$
(2.5)

By Lemma 2.13, we have

$$\begin{aligned} (K_0+\alpha I_{N(F)})^{-1}-\alpha ^{-1} I_{N(F)}=-\alpha ^{-1}K_0(K_0+\alpha I_{N(F)})^{-1}, \end{aligned}$$

and hence

$$\begin{aligned} (K_0+\alpha I_{N(F)})^{-1}=\alpha ^{-1} I_{N(F)}-\alpha ^{-1}K_0(K_0+\alpha I_{N(F)})^{-1}. \end{aligned}$$

With similar arguments, we can obtain \((\alpha I_{R(F)}-F_0)^{-1}=\alpha ^{-1}I_{R(F)}+\alpha ^{-1}F_0(\alpha I_{R(F)}-F_0)^{-1}\). Substituting these quantities in equation (2.5), we obtain the representation of \(T^{-1}\) as in (3). \(\square \)

Remark 2.15

Let

$$\begin{aligned} \beta&=\alpha ^{-1},\\ K_1&= \left( \begin{array}{cc} \alpha ^{-1}K_0(K_0+\alpha I_{N(F)})^{-1} &{} 0 \\ 0 &{} 0 \\ \end{array}\right) \end{aligned}$$

and

$$\begin{aligned} F_1&=\left( \begin{array}{cc} 0&{} 0 \\ 0 &{} \alpha ^{-1}F_0(\alpha I_{R(F)}-F_0)^{-1} \\ \end{array} \right) . \end{aligned}$$

Then \(T^{-1}=\beta I-K_1+F_1\). Note that \(\Vert K_1\Vert \le \beta \), since \(\Vert K_0(\alpha I_{N(F)}+K_0)^{-1}\Vert \le 1\). Clearly, by definition, \(K_1F_1=0\). This is exactly the structure of absolutely minimum attaining operators (shortly \(\mathcal {AM}\)-operators) in the case when T is positive and one-to-one. We refer [6] for more details of the structure of these operators. We recall that \(A\in {\mathcal {B}}(H_1,H_2)\) is said to be minimum attaining if there exists \(x_0\in S_{H_1}\) such that \(\Vert Ax_0\Vert =m(A)\) and absolutely minimum attaining if \(A|_{M}\) is minimum attaining for each non zero closed subspace M of \(H_1\).

PROPOSITION 2.16

Let \(T\in {\mathcal {B}}(H)\) satisfy condition (2) of Theorem 2.5. Then, with respect the pair of subspace \((N(K), N(K)^{\bot })\), T has the following decomposition:

$$\begin{aligned} T=\left( \begin{array}{cc} \alpha I_{N(K)}-F_0 &{} 0 \\ 0 &{} K_0+\alpha I_{N(K)^{\bot }} \\ \end{array} \right) , \end{aligned}$$

where \(F_0=F|_{N(K)}\) and \(K_0=K|_{N(K)^{\bot }}\).

Proof

First, we show that N(K) is a reducing subspace for T. Since \(T\ge 0\), it suffices to show that N(K) is invariant under T. For this, let \(x\in N(K)\). Then \(Tx=(\alpha I-F)(x)\) and \(K(Tx)=(\alpha I-F)(Kx)=0\). This proves the claim. Next, if \(x\in N(K)\), then \(Tx=(\alpha I-F)(x)\). That is, \(T|_{N(K)}=\alpha I_{N(K)}-F|_{N(K)}\).

If \(y\in N(K)^{\bot }=\overline{R(K)}\), then there exists a sequence \((x_n)\subset H\) such that \(y= \lim _{n\rightarrow \infty }Kx_n\). So \(Fy=\lim _{n\rightarrow \infty }FKx_n=0\). Thus we have \(Ty=Ky+\alpha y\). So \(T|_{N(K)^{\bot }}=K_{N(K)^{\bot }}+\alpha I_{N(K)^{\bot }}\). \(\square \)

3 Self-adjoint and normal \({\mathcal {AN}}\)-operators

In this section, we first discuss the structure of self-adjoint \({\mathcal {AN}}\)-operators. Later, we extend this to the case of normal operators.

Theorem 3.1

Let \(T=T^*\in {{\mathcal {AN}}}(H)\). Then there exists an orthonormal basis consisting of eigenvectors of T.

Proof

The proof follows along the similar lines of [10, Theorem 3.1]. For the sake of completeness, we provide the details here. Let \(\mathcal B=\{x_\alpha : \alpha \in I\}\) be the maximal set of orthonormal eigenvectors of T. This set is non empty as \(T=T^*\in {\mathcal {AN}}(H)\). Let \(M=\overline{\text{ span }}\{x_\alpha : \alpha \in I\}\). Then we claim that \(M=H\). If not, \(M^{\bot }\) is a proper non-zero closed subspace of H and it is invariant under T. Since \(T=T^* \in {\mathcal {AN}}(H)\), we have either \(||T|M^{\bot }||\) or \(-||T|M^{\bot }||\) is an eigenvalue for \(T|M^{\bot }\). Hence there is a non-zero vector, say \(x_0\) in \(M^{\bot }\), such that \(Tx_0=\pm ||T|M^{\bot }|| x_0.\) Hence \(x_{0}\in M\). Since \(M\cap M^{\bot }={\{0}\},\) a contradiction. \(\square \)

PROPOSITION 3.2

Let \(T=T^*\in {{\mathcal {AN}}}(H)\). Then the following holds:

  1. (1)

    T can have atmost two eigenvalues with infinite multiplicity. Moreover, if \(\alpha \) and \(\beta \) are such eigenvalues, then \(\alpha =\pm \beta \),

  2. (2)

    if T has an eigenvalue \(\alpha \) with infinite multiplicity and \(\beta \) is a limit point of \(\sigma (T)\), then \(\alpha =\pm \beta \),

  3. (3)

    \(\sigma (T)\) can have atmost two limit points. If \(\alpha \) and \(\beta \) are such points, then \(\alpha =\pm \beta \).

Proof

Proof of (1). Let \(\alpha _j\in \sigma _p(T)\) be such that \(N(T-\alpha _jI)\) is infinite dimensional for each \(j=1,2,3\). Then \(\alpha _j^2\in \sigma _p(T^2)\) and we have \(N(T-\alpha _jI)\subseteq N(T^2-\alpha _j^2I)\) for each \(j=1,2,3\). Since \(T^2\in {\mathcal {AN}}(H)\) and positive, by (3) of Theorem 2.2, it follows that \(\alpha _1^2=\alpha _2^2=\alpha _3^2\). Thus \(\alpha _1=\pm \alpha _2=\pm \alpha _3\).

Proof of (2). Let \(\alpha \in \sigma _p(T)\) with infinite multiplicity and \(\beta \in \sigma (T)\), which is a limit point. Since \(\sigma (T^2)={\{\lambda ^2:\lambda \in \sigma (T)}\}\), it follows that \(\alpha ^2\) is an eigenvalue of \(T^2\) with infinite multiplicity as \(N(T-\alpha I)\subseteq N(T^2-\alpha ^2I)\) and \(\beta ^2\) is a limit point \(\sigma (T^2)\). Since \(T^2\in {\mathcal {AN}}(H)\) is positive, by (4) of Theorem (2.2), \(\alpha ^2=\beta ^2\). Thus \(\alpha =\pm \beta \).

Proof of (3). Let \(\alpha ,\beta \in \sigma (T)\) be limit points of \(\sigma (T)\). Then \(\alpha ^2,\beta ^2\in \sigma (T^2)\) are limit points of \(\sigma (T^2)\) and since \(T^2\in {\mathcal {AN}}(H)\) and positive, by Theorem 2.2(2), \(\alpha ^2=\beta ^2\), concluding that \(\alpha =\pm \beta \). By arguing as in Proof of (1), we can show that there are at most two limit points for the spectrum. \(\square \)

The following decomposition of a self-adjoint operator is used in the sequel:

Let \(T=T^*\in {\mathcal {B}}(H)\) and have the polar decomposition \(T=V|T|\). Let \(H_0=N(T),\; H_{+}=N(I-V)\) and \(H_{-}=N(I+V)\). Then \(H=H_0\oplus H_{+}\oplus H_{-}\). All these subspaces are invariant under T. Let \(T_0=T|_{N(T)},\; T_{+}=T|_{H_{+}}\) and \(T_ {-}=T|_{H_{-}}\). Then \(T=T_0\oplus T_{+}\oplus T_{-}\). Further more, \(T_{+}\) is strictly positive, \(T_{-}\) is strictly negative and \(T_0=0\) if \(N(T)\ne {\{0}\}\). Let \(P_0=P_{N(T)},\; P_{\pm }=P_{H_{\pm }}\). Then \(P_0=I-V^2\) and \(P_{\pm }=\frac{1}{2}(V^2\pm V)\). Thus \(V=P_{+}-P_{-}\). Moreover, \(|T|=T_{+}\oplus (-T_{-})\oplus T_0\). For details, see [9, Example 7.1, p. 139]. Note that the operators \(T_{+}\) and \(T_{-}\) are different than those used in Theorem 2.5.

Theorem 3.3

Let \(T\in {{\mathcal {AN}}}(H)\) be self-adjoint with the polar decomposition \(T=V|T|\). Then

  1. (1)

    the operator T has the representation

    $$\begin{aligned} T=K-F+\alpha V, \end{aligned}$$

    where \(K\in {\mathcal {K}}(H),\; F\in {\mathcal {F}}(H)\) are self-adjoint with \(KF=0=FK,\; \alpha \ge 0\) and \(F^2\le \alpha ^2I\),

  2. (2)

    if T is not a compact operator, then \(V\in {\mathcal {AN}}(H)\),

  3. (3)

    \(K^2+2\alpha \text {Re}(VK)\ge 0\).

Proof

Proof of (1). Since T is self-adjoint, by considering \(H_0,H_{\pm }\) and \(T_0,T_{\pm }\) as in the earlier discussion, we can write \(H=H_{0}\oplus H_{+}\oplus H_{-}\) and \(T=T_0\oplus T_{+}\oplus T_{-}\). Note that \(T_0=0\) if \(H_{0}\ne {\{0}\}\). Since \(H_{\pm }\) reduces T, we have \(T_{\pm }\in \mathcal B(H_{\pm })\). As \(T\in {{\mathcal {AN}}}(H)\), we have that \(T_{\pm }\in {\mathcal {AN}}(H_{\pm })\). Hence by Theorem 2.5, we have \(T_{+}=K_{+}-F_{+}+\alpha I_{H_{+}}\) such that \(K_{+}\) is a positive compact operator, \(F_{+}\) is a finite rank positive operator with the property that \(K_{+}F_{+}=0\) and \(F_{+}\le \alpha I_{H_{+}} \). As \(T_{+}\) is strictly positive, \(\alpha >0\), in fact, \(\alpha =m_e(T_{+})\).

Similarly, \(T_{-}\in {{\mathcal {AN}}}(H_{-})\) and strictly negative. Hence there exists a triple \((K_{-},F_{-},\beta )\) such that \(-T_{-}=K_{-}-F_{-}+\beta I_{H_{-}}\), where \(K_{-}\in \mathcal K(H_{-})\) is positive, \(F_{-}\in {\mathcal {F}}(H_{-})\) is positive with \(K_{-}F_{-}=0,\; F_{-}\le \beta I_{H_{-}}\) and \(\beta >0\), in fact, \(\beta =m_e(T_{-})\). Hence we can write \(T_{-}=-K_{-}+F_{-}-\beta I_{H_{-}}\). So, we have

$$\begin{aligned} T&=\left( \begin{array}{ccc} T_{+} &{} 0 &{} 0 \\ 0 &{} T_{-} &{} 0 \\ 0 &{} 0 &{} T_0 \\ \end{array} \right) \\&=\left( \begin{array}{ccc} K_{+}-F_{+}+\alpha I_{H_{+}}&{} 0 &{} 0 \\ 0 &{} -K_{-}+F_{-}-\beta I_{H_{-}} &{} 0 \\ 0 &{} 0 &{} 0 \\ \end{array} \right) \\&=\left( \begin{array}{ccc} K_{+} &{}0 &{}0 \\ 0 &{} -K_{-} &{} 0 \\ 0 &{} 0 &{} 0 \\ \end{array} \right) -\left( \begin{array}{ccc} F_{+} &{} 0 &{} 0 \\ 0 &{} -F_{-} &{} 0 \\ 0 &{} 0 &{} 0 \\ \end{array} \right) +\left( \begin{array}{ccc} \alpha I_{H_{-}} &{}0 &{} 0 \\ 0 &{} -\beta I_{H_{-}} &{}0 \\ 0 &{}0 &{}0 \\ \end{array} \right) . \end{aligned}$$

Note that \(|T|=T_{+}\oplus (-T_{-})\oplus T_0\) (see [9, Example 7.1, p. 139] for details). That is,

$$\begin{aligned} |T|&=\left( \begin{array}{ccc} K_{+} &{}0 &{}0 \\ 0 &{} K_{-} &{} 0 \\ 0 &{} 0 &{} 0 \\ \end{array} \right) -\left( \begin{array}{ccc} F_{+} &{} 0 &{} 0 \\ 0 &{}F_{-} &{} 0 \\ 0 &{} 0 &{} 0 \\ \end{array} \right) +\left( \begin{array}{ccc} \alpha I_{H_{-}} &{}0 &{} 0 \\ 0 &{} -\beta I_{H_{-}} &{}0 \\ 0 &{}0 &{}0 \\ \end{array} \right) \\&=\left( \begin{array}{ccc} K_{+} &{}0 &{}0 \\ 0 &{} K_{-} &{} 0 \\ 0 &{} 0 &{} 0 \\ \end{array} \right) -\left( \begin{array}{ccc} F_{+} &{} 0 &{} 0 \\ 0 &{}F_{-} &{} 0 \\ 0 &{} 0 &{} \alpha I_{H_0} \\ \end{array} \right) +\left( \begin{array}{ccc} \alpha I_{H_{-}} &{}0 &{} 0 \\ 0 &{} -\beta I_{H_{-}} &{}0 \\ 0 &{}0 &{}\alpha I_{H_0}\\ \end{array} \right) . \end{aligned}$$

Let \(K=\left( \begin{array}{ccc} K_{+} &{}0 &{}0 \\ 0 &{} -K_{-} &{} 0 \\ 0 &{} 0 &{} 0 \\ \end{array} \right) \) and \(F=\left( \begin{array}{ccc} F_{+} &{} 0 &{} 0 \\ 0 &{}-F_{-} &{} 0 \\ 0 &{} 0 &{} \alpha I_{H_0} \\ \end{array} \right) \).

Since \(|T|\in {\mathcal {AN}}(H)\) and positive, by the uniqueness of the representation, we get that \(\alpha =\beta \).

Now let, \(V=\left( \begin{array}{ccc} I_{H_{+}} &{}0 &{} 0 \\ 0 &{} - I_{H_{-}} &{}0 \\ 0 &{}0 &{}0 \\ \end{array} \right) \). Then \(T=K-F+\alpha V\) and K and F satisfy the stated properties.

If T is one-to-one, then \(N(T)={\{0}\}\). In this case, \(H=H_{+}\oplus H_{-}\) and \(T=T_{+}\oplus T_{-}\). Arguing as above, we can obtain the representation for \(T=K-F+\alpha V\), where

$$\begin{aligned} K=\left( \begin{array}{cc} K_{+} &{} 0 \\ 0 &{} -K_{-} \\ \end{array} \right) ,\; F=\left( \begin{array}{cc} F_{+} &{} 0 \\ 0 &{} -F_{-} \\ \end{array} \right) ,\; V=\left( \begin{array}{cc} \alpha I_{H_+} &{} 0 \\ 0 &{} -\alpha I_{H_{-}} \\ \end{array} \right) . \end{aligned}$$

Proof of (2). Note that if \(\alpha =0\), then T is compact. If \(\alpha >0\) and V is a finite rank operator, then also T is compact. Since we assumed that T is not compact, it must be the case that \(\alpha >0\) and R(V) is infinite dimensional. But by Proposition 2.8, \(N(T)=N(V)\) is finite dimensional. So the conclusion follows by [3, Proposition 3.14].

Proof of (3). First note that since \(T=T^*\), it follows that \(V=V^*\). As \(VK=KV\), KV is self-adjoint. Hence \(K^2+2\,\mathrm{Re}(V^*K)=K^2+2VK\). Thus

$$\begin{aligned} K^2+2VK=\left( \begin{array}{ccc} K_{+}^2+2K_{+} &{}0 &{}0 \\ 0 &{} K_{-}^2-2K_{-} &{} 0 \\ 0 &{} 0 &{} 0 \\ \end{array} \right) . \end{aligned}$$

Since the (1, 1) entry of the above matrix is positive, to get the conclusion, it suffices to prove that the (2, 2) entry is positive. Clearly, \(K_{-}^2-2K_{-}\) is self-adjoint. Next, we show that \(\sigma (K_{-}^2-2K_{-})\) is positive. Let \(\lambda \in \sigma (K_{-})\). Then \(\lambda \le 0\) and \(\lambda ^2-2\lambda \in \sigma (K_{-}^2-2K_{-})\). But \(\lambda ^2-2\lambda =\lambda (\lambda -2)\ge 0\). Hence \(K_{-}^2-2K_{-}\) is positive. \(\square \)

COROLLARY 3.4

Let \(T=T^*\in {\mathcal {AN}}(H)\). Then \(\sigma (T)\) is countable.

Proof

Since \(T=T_{+}\oplus T_{-}\oplus T_0\), all these operators \(T_{+},T_{-}\) and \(T_0\) are \({\mathcal {AN}}\) operators. We know that \(\sigma (T_{+}),\sigma (T_0)\) are countable, as they are positive. Also, \(-T_{-}\) is a positive \({{\mathcal {AN}}}\)-operator and hence \(\sigma (T_{-})\) is countable. Hence, we can conclude that \(\sigma (T)=\sigma (T_{+})\cup \sigma (T_{-})\cup \sigma (T_0)\) is countable. \(\square \)

Next, we can get the structure of normal \({\mathcal {AN}}\)-operators. Here we use a different approach to the one used in Theorem 3.3.

PROPOSITION 3.5

Let \(T\in {\mathcal {AN}}(H)\) be normal with the polar decomposition \(T=V|T|\). Then there exists a compact normal operator K, a finite rank normal operator \(F\in {\mathcal {B}}(H)\) and \(\alpha \ge 0\) such that

  1. (1)

    T has the representation

    $$\begin{aligned} T=K-F+\alpha V \end{aligned}$$
    (3.1)

    with \(KF=0\) and \(F^*F\le \alpha ^2I\),

  2. (2)

    \(K^*K+2\alpha \text {Re}(V^*K)\ge 0\),

  3. (3)

    VKF commutes mutually,

  4. (4)

    if \(\alpha >0\), then \(V\in {{\mathcal {AN}}}(H)\).

Proof

Proof of (1). Since T is normal, \(|T|=|T^*|\). Hence \(|T^*|^2=TT^*=V|T|^2V^*\). That is \(V^*|T^*|^2=V^*V|T|^2V^*=|T|^2V^*\) or \(V|T|^2=|T|^2V\). By the square root property, it follows that \(V|T|=|T|V\).

Since \(|T|\in {\mathcal {AN}}(H)\), we have \(|T|=K_1-F_1+\alpha I\), where \(K_1\in {\mathcal {K}}(H)\) is positive, \(F_1\in {\mathcal {F}}(H)\) is positive and \(F_1\le \alpha I\).

Next, we show that V is normal. We have \(N(T^*)=N(T)=N(V)\). Hence

$$\begin{aligned} V^*V=P_{N(V)^{\bot }}=P_{N(T)^{\bot }}=P_{N(T^*)^{\bot }}=P_{\overline{R(T)}}=P_{R(V)}=VV^*. \end{aligned}$$

So \(T=K-F+\alpha V\), where \(K=VK_1\) and \(F=VF_1\). Next, we show that K and F are normal. As T is normal, V commutes with |T|, we have

$$\begin{aligned} V(K_1-F_1)=(K_1-F_1)V. \end{aligned}$$
(3.2)

Since V commutes with \(K_1-F_1\), it also commutes with \((K_1-F_1)^2\). But, \((K_1-F_1)^2=K_1^2+F_1^2=(K_1+F_1)^2\). With this, we can conclude that \(V(K_1+F_1)^2=(K_1+F_1)^2V\). Hence,

$$\begin{aligned} V(K_1+F_1)=(K_1+F_1)V. \end{aligned}$$
(3.3)

Thus by equations (3.2) and (3.3), we can conclude that \(VK_1=K_1V\) and \(VF_1=F_1V\). By the Fuglede’s theorem we can conclude that \(V^*K_1=K_1V^*\) and \(V^*F_1=F_1V^*\). Next,

$$\begin{aligned} K^*K=K_1V^*VK_1=K_1VV^*K_1 =VK_1V^*K_1=VK_1K_1V^*=KK^*. \end{aligned}$$

With similar arguments we can show that F is normal.

Next, we show that \(KF=0\). Since V commutes with \(K_1\) and \(F_1\), we have \(KF=VK_1VF_1=V^2K_1F_1=0\).

Finally, \(F^*F=F_1V^*VF_1\le \Vert V\Vert ^2F_1^2\le \alpha ^2I\).

Proof of (2). Using the relations \(VK_1=K_1V\) and \(V^*K_1=K_1V^*\), we get

$$\begin{aligned} K^*K+\alpha (V^*K+K^*V)&=K_1V^*VK_1+\alpha (V^*VK_1+K_1V^*V)\\&=V^*V(K_1^2+2\alpha K_1) \\&=P_{N(V)^{\bot }} (K_1^2+2\alpha K_1)\\&=K_1^2+2\alpha K_1\\&\ge 0. \end{aligned}$$

In the fourth step of the above equations, we have used the fact that \(P_{N(V)^{\bot }}K_1=P_{R(V)}K_1=P_{R(|T|)}K_1=K_1\).

Proof of (3). We have \(VK=VVK_1=VK_1V=KV\) and \(VF=VVF_1=VF_1V=FV\). Also, \(KF=0=FK\).

Proof of (4). Note that by applying Proposition 2.8(2) to |T|, we can conclude that \(N(|T|)=N(T)=N(V)\) is finite dimensional. Now the conclusion follows by [3, Proposition 3.14]. \(\square \)

COROLLARY 3.6

Let \(T\in {\mathcal {B}}(H)\) be normal. Then \(T\in {\mathcal {AN}}(H)\) if and only if \(T^*\in {\mathcal {AN}}(H)\).

Proof

We know that \(T\in {\mathcal {AN}}(H)\) if and only if \(T^*T\in {\mathcal {AN}}(H)\), by Corollary 2.11. Since \(T^*T=TT^*\), by Corollary 2.11 again, it follows that \(TT^*\in {\mathcal {AN}}(H)\) if and only if \(T^*\in {\mathcal {AN}}(H)\). \(\square \)

4 General case

In this section, we prove the structure of absolutely norm attaining operators defined between two different Hilbert spaces.

Theorem 4.1

Let \(T\in {{\mathcal {AN}}}(H_1,H_2)\) with the polar decomposition \(T=V|T|\). Then

$$\begin{aligned} T=K-F+\alpha V, \end{aligned}$$

where \(K\in {\mathcal {K}}(H_1,H_2), \; F\in {\mathcal {F}}(H_1,H_2)\) such that \(K^*F=0=KF^*\) and \(\alpha ^2I\ge F^*F \).

Proof

Since \(|T|\in {\mathcal {AN}}(H_1)\) and positive, we have by Theorem 2.5, \(|T|=K_1-F_1+\alpha I\), where the triple \((K_1,F_1,\alpha )\) satisfy conditions in Theorem 2.5(2). Now, \(T=K-F+\alpha V\), where \(K=VK_1,\; F=VF_1\). Clearly,

$$\begin{aligned} K^*F=K_1V^*VF_1=K_1P_{N(V)^{\bot }}F_1&=K_1(I-P_{N(V)})F_1\\&=K_1F_1-K_1P_{N(V)}F_1\\&=0 \;(\text {since }\; N(V)=N(|T|)\subseteq N(K_1)) . \end{aligned}$$

Also, clearly, \(KF^*=VK_1F_1V^*=0\).

Finally, \(F^*F=F_1V^*VF_1\le \Vert V^*V\Vert F_1^2\le F_1^2 \le \alpha ^2 I\). \(\square \)

PROPOSITION 4.2

Let \(T\in {\mathcal {B}}(H)\) and \(U\in {\mathcal {B}}(H)\) be unitary such that \(T^*=U^*TU\). Then \(T\in {{\mathcal {AN}}}(H)\) if and only if \(T^*\in {{\mathcal {AN}}}(H)\).

Proof

This follows by [3, Theorem 3.5]. \(\square \)

Next, we discuss a possible converse in the general case.

Theorem 4.3

Let \(K\in {\mathcal {K}}(H_1,H_2),\; F\in {\mathcal {F}}(H_1,H_2)\), \(\alpha \ge 0\) and \(V\in {\mathcal {B}}(H_1,H_2)\) be a partial isometry. Further, assume that

  1. (1)

    \(V\in {\mathcal {AN}}(H_1,H_2)\),

  2. (2)

    \( K^*K+\alpha (V^*K+K^*V)\ge 0\).

Then \(T:=K-F+\alpha V \in {{\mathcal {AN}}}(H_1,H_2)\).

Proof

If \(\alpha =0\), then \(T\in {\mathcal {K}}(H_1,H_2)\). Hence \(T\in {{\mathcal {AN}}}(H_1,H_2)\). Next assume that \(\alpha >0\). We prove this case by showing \(T^*T\in {\mathcal {AN}}(H_1)\). By a simple calculation, we can get \( T^*T={\mathcal {K}}-{\mathcal {F}} +\alpha ^2 P_{N(V)^{\bot }}\), where

$$\begin{aligned} {\mathcal {K}}= K^*K+\alpha (V^*K+K^*V),\quad {\mathcal {F}}= F^*F-F^*K-K^*F-\alpha (V^*F+F^*V). \end{aligned}$$

Since \(V\in {\mathcal {AN}}(H_1,H_2)\), either N(V) or \(N(V)^{\bot }\) is finite dimensional by [3, Proposition 3.14]. If \(N(V)^{\bot }\) is finite dimensional, then \(T^*T\in {\mathcal {K}}(H_1)\). Hence \(T\in {\mathcal {K}}(H_1,H_2)\).

If N(V) is finite dimensional, then \(T^*T={\mathcal {K}}-(\mathcal F-\alpha ^2 P_{N(V)})+\alpha ^2I\). Note that the operator \(\mathcal F-\alpha ^2 P_{N(V)}\) is a finite rank self-adjoint operator. Hence \(T^*T\in {\mathcal {AN}}(H_1)\), by Theorem 2.1. Now the conclusion follows by Corollary 2.11. \(\square \)

COROLLARY 4.4

Suppose that \(K\in {\mathcal {K}}(H)\), \( F\in {\mathcal {F}}(H)\) are normal and \(V\in {\mathcal {B}}(H)\) is a normal partial isometry such that VFK commute mutually. Let \(\alpha \ge 0\). Then

  1. (1)

    \(T:=K-F+\alpha V\) is normal,

  2. (2)

    if \(K^*K+2\alpha V^*K \ge 0\) and \(V\in {{\mathcal {AN}}}(H)\), then \(T\in {{\mathcal {AN}}}(H)\).

Proof

Proof of (1). We observe that if A and B are commuting normal operators, then \(A+B\) is normal (see [12, p. 342, Exercise 12] for details). By this observation, it follows that T is normal.

Proof of (2). Since \(VK=KV\), by Fuglede’s theorem [12, p. 315], \(V^*K=KV^*\). With this observation and Theorem 4.3, the conclusion follows. \(\square \)

COROLLARY 4.5

Suppose that \(K\in {\mathcal {K}}(H)\), \( F\in {\mathcal {F}}(H)\) are self-adjoint and \(V\in {\mathcal {B}}(H)\) is a self-adjoint, partial isometry and \(\alpha \ge 0\) such that

  1. (a)

    \(V\in {\mathcal {AN}}(H)\),

  2. (b)

    \(K^2+2\alpha (VK)\ge 0\).

Then \(T:=K-F+\alpha V\) is self-adjoint \({\mathcal {AN}}\)-operator.

Proof

The proof directly follows by Theorem 4.3. \(\square \)

DEFINITION 4.6

[4, p. 349]

Let \(T\in {\mathcal {B}}(H_1,H_2)\). Then T is called left-semi-Fredholm if there exists \(B\in \mathcal B(H_2,H_1)\) and \(K\in {\mathcal {K}}(H_1)\) such that \(BT=K+I\), and right-semi-Fredholm if there exists \(A\in \mathcal B(H_2,H_1)\) and \(K'\in {\mathcal {K}}(H_2)\) such that \(TA=K'+I\).

If T is both left-semi-Fredholm and right-semi-Fredholm, then T is called Fredholm.

Remark 4.7

Note that T is left semi-Fredholm if and only if \(T^*\) is right semi-Fredholm (see [4, section 2, p. 349] for details).

COROLLARY 4.8

Let \(T\in {\mathcal {AN}}(H_1,H_2)\) but not compact. Then T is left-semi-Fredholm.

Proof

Let \(T=V|T|\) be the polar decomposition of T. Then \(|T|=V^*T\). As \(|T|\in {\mathcal {AN}}(H_1)\), by Theorem 2.5, there exists a triple \((K,F,\alpha )\) satisfying conditions in Theorem 2.5, such that \(V^*T=K-F+\alpha I\). Let \(K^{'}=K-F\). Then \(V^*T=K^{'}+\alpha I\). By Definition 4.6, it follows that T is left-semi-Fredholm. \(\square \)