1 Introduction

Many problems arising in different areas of mathematics, such as optimization, variational analysis, signal recovery, etc., can be modeled by the fixed point problem

$$\begin{aligned}\text {Find} \quad x\in K \quad \text {such that} \quad x=Tx,\end{aligned}$$

where \(T:K\rightarrow K\) is a mapping and K is a nonempty subset of a Hilbert space H. The solutions to this problem are called fixed points of the mapping T. We denote the set of fixed points of T by

$$\begin{aligned}F(T):=\{x\in K \mid x=Tx\}.\end{aligned}$$

A mapping \(T:K\rightarrow K\) is said to be nonexpansive if

$$\begin{aligned}\Vert Tx-Ty\Vert \le \Vert x-y\Vert ,\end{aligned}$$

for all \(x,y\in K\).

A significant body of work on iteration methods for fixed points problems of nonexpansive mappings has accumulated in the literature (for example, see [1, 3, 5, 9, 11]). In particular, one of the most famous fixed point methods is the Krasnosel’skiĭ–Mann iteration from [14, 17, 20] that starts at some given point \(x_0\in K\) and uses the recursion

$$\begin{aligned} x_{n+1}=(1-\lambda _n)x_n+\lambda _nTx_n, \end{aligned}$$
(1.1)

for some suitably chosen scalars \(\lambda _n\in [0,1]\). The iterative sequence \(\{x_n\}\) converges weakly to a fixed point of T provided that F(T) is nonempty and \(\lambda _n\) satisfies

$$\begin{aligned}\sum _{n=0}^\infty \lambda _n(1-\lambda _n)=\infty .\end{aligned}$$

Another famous fixed point method is the Halpern iteration, which was first introduced in 1967 by Halpern [10] in the framework of Hilbert spaces. The recursion formula of the Halpern iteration is as follows: \(x_0\in K\),

$$\begin{aligned} x_{n+1}=\delta _n u+(1-\delta _n)Tx_n, \end{aligned}$$
(1.2)

where \(u\in K\) is an arbitrary (but fixed) element in K, and \(\delta _n\in [0,1]\). Halpern proved the strong convergence of \(\{x_n\}\) to a fixed point of T where \(\delta _n:=n^{-a}, a\in (0,1)\). In 1977, Lions [15] improved Halpern’s result by proving strong convergence of \(\{x_n\}\) to a fixed point of T if the real sequence \(\{\delta _n\}\) satisfies the following conditions:

  1. (A1)

    \(\lim _{n\rightarrow \infty }\delta _n=0\) and \(\sum _{n=0}^\infty \delta _n=\infty \);

  2. (A2)

    \(\lim _{n\rightarrow \infty }\frac{|\delta _n-\delta _{n-1}|}{\delta _n^2}=0\).

It is easy to observe that both Halpern’s and Lion’s conditions on the sequence \(\{\delta _n\}\) excluded the canonical choice \(\delta _n=\frac{1}{n+1}\). To overcome this, Wittmann [28] proved the strong convergence of \(\{x_n\}\) to a fixed point of T if \(\{\delta _n\}\) satisfies condition (A1) and the following condition (A3):

  1. (A3)

    \(\sum _{n=0}^\infty |\delta _{n+1}-\delta _n|<\infty \).

In 2002, Xu [21, 29, 30] suggested the following condition (A4) instead of the conditions (A2) or (A3):

  1. (A4)

    \(\lim _{n\rightarrow \infty }\frac{|\delta _n-\delta _{n-1}|}{\delta _n}=0\).

He proved the strong convergence of \(\{x_n\}\) using the control conditions (A1) and (A4). Xu also showed that condition (A2) and condition (A4) are not comparable.

Chidume and Chidume [7] and Suzuki [24] proposed the following simpler modification of the Halpern iteration method: \(x_0\in K\):

$$\begin{aligned} x_{n+1}=\delta _n u+(1-\delta _n)\big ((1-\alpha )x_n+\alpha Tx_n\big ), \end{aligned}$$
(1.3)

where \(u\in K\) is an arbitrary (but fixed) element in K, and \(\alpha \in (0,1)\) is a constant. They proved the strong convergence of the iteration (1.3) only under condition (A1). Kim and Xu [13] proposed the following iteration method: \(x_0\in K\):

$$\begin{aligned} \left\{ \begin{array}{ll} y_n=\beta _n x_n+(1-\beta _n)Tx_n, \\ x_{n+1}=\alpha _n u+(1-\alpha _n)y_n, \end{array} \right. \end{aligned}$$
(1.4)

where \(\{\alpha _n\}\) and \(\{\beta _n\}\) are two sequences in (0, 1). Assume that the following conditions are satisfied:

  1. (B1)

    \(\lim _{n\rightarrow \infty }\alpha _n= 0, \sum _{n=0}^\infty \alpha _n=\infty \) and \(\sum _{n=0}^\infty |\alpha _{n+1}-\alpha _n|<\infty \);

  2. (B2)

    \(\lim _{n\rightarrow \infty }\beta _n= 0, \sum _{n=0}^\infty \beta _n=\infty \) and \(\sum _{n=0}^\infty |\beta _{n+1}-\beta _n|<\infty \).

Then, \(\{x_n\}\) defined by (1.4) converges strongly to a fixed point of T. Yao [33] also proved that the sequence \(\{x_n\}\) defined by (1.4) converges strongly to a fixed point of T provided that the control sequences \(\{\alpha _n\}\) and \(\{\beta _n\}\) satisfy the conditions

  1. (B3)

    \(\lim _{n\rightarrow \infty }\alpha _n=0\) and \(\sum _{n=0}^\infty \alpha _n=\infty \);

  2. (B4)

    \(0<\liminf _{n\rightarrow \infty }\beta _n\le \limsup _{n\rightarrow \infty }\beta _n<1\).

Song and Li [23] proposed the slightly different iteration process: \(x_0\in K\)

$$\begin{aligned} x_{n+1}=\beta _n(\alpha _n u+(1-\alpha _n)x_n)+(1-\beta _n)Tx_n. \end{aligned}$$
(1.5)

They proved that the conditions (B3) and (B4) are sufficient to guarantee the strong convergence of the sequence \(\{x_n\}\) generated by the algorithm (1.5).

Inspired and motivated by the above studies, Kanzow and Shehu [12] introduced a new inexact generalized Halpern iteration method. To be more precise, they consider the following iteration process: \(x_0\in H\):

$$\begin{aligned} x_{n+1}=\delta _n u+\alpha _n x_n+\beta _n Tx_n+r_n, \end{aligned}$$
(1.6)

where \(T:H\rightarrow K\) is a nonexpansive mapping, \(u\in K\) denotes a fixed element, \(r_n\) represents the residual, and \(\delta _n, \alpha _n, \beta _n\in [0,1]\) are chosen, such that \(\delta _n+\alpha _n+\beta _n\le 1\). Clearly, the iterative sequence (1.6) is a natural generalization of the iterations (1.3), (1.4) and (1.5). They proved that the sequence \(\{x_n\}\) generated by (1.6) converges strongly to a point in F(T), which is the projection of u to F(T), under the following conditions:

  1. (C1)

    \(\lim _{n\rightarrow \infty }\delta _n=0\) and \(\sum _{n=0}^\infty \delta _n=\infty \);

  2. (C2)

    \(\sum _{n=0}^\infty (1-\delta _n-\alpha _n-\beta _n)<\infty \);

  3. (C3)

    \(\sum _{n=0}^\infty \Vert r_n\Vert <\infty \);

  4. (C4)

    \(\liminf _{n\rightarrow \infty }\alpha _n\beta _n>0\).

It is easy to verify that Kanzow’s conditions on the sequences \(\{\delta _n\}, \{\alpha _n\}, \{\beta _n\}\) and \(\{r_n\}\) exclude the canonical choices \(\delta _n=\frac{1}{2(n+1)}, \alpha _n=\frac{1}{2(n+1)}, \beta _n=1-\frac{1}{n+1}\) and \(r_n\equiv 0\). In this paper, we will prove that the sequence \(\{x_n\}\) generated by (1.6) converges strongly to a point in F(T) under this choices. On the other hand, since \(\alpha _n\equiv 0\) does not satisfy condition (C4), the iteration (1.6) does not include (1.2) as a special case. To overcome these shortcomings, we give certain different control conditions on the sequences \(\{\delta _n\}, \{\alpha _n\}, \{\beta _n\}\) and \(\{r_n\}\) so that \(\{x_n\}\) generated by (1.6) converges strongly to a fixed point of T. Moreover, we give the convergence rate of the generalized Halpern iteration under specially selected parameters, which satisfy our conditions, but do not satisfy Kanzow’s conditions.

2 Preliminaries

Throughout this paper, we denote by H a real Hilbert space with inner product \(\langle \cdot , \cdot \rangle \) and norm \(\Vert \cdot \Vert \) and by I the identity operator on H. The strong (weak) convergence of a sequence \(\{x_n\}\) to x is denoted by \(x_n\rightarrow x\) (\(x_n\rightharpoonup x\)), respectively.

Definition 2.1

[2] Let D be a nonempty subset of H and let \(T:D\rightarrow H\). Then, T is

  1. (1)

    Nonexpansive if

    $$\begin{aligned}\Vert Tx-Ty\Vert \le \Vert x-y\Vert , \quad \forall x, y\in D;\end{aligned}$$
  2. (2)

    Firmly nonexpansive if

    $$\begin{aligned} \Vert Tx-Ty\Vert ^2\le \Vert x-y\Vert ^2-\Vert (I-T)x-(I-T)y\Vert ^2, \quad \forall x,y\in D; \end{aligned}$$
  3. (3)

    \(\nu \)-inverse strongly monotone if there is \(\nu >0\) such that

    $$\begin{aligned}\langle Tx-Ty, x-y\rangle \ge \nu \Vert Tx-Ty\Vert ^2, \quad \forall x,y\in D.\end{aligned}$$

For any \(x\in H\), the projection to a nonempty closed convex subset C is defined as

$$\begin{aligned}P_C x= \text {argmin}\{\Vert y-x\Vert \mid y\in C\}.\end{aligned}$$

The projection \(P_C\) has the following well-known properties.

Lemma 2.2

[2, 9] Let \(C\subseteq H\) be a nonempty closed convex subset. Then, for all \(x,y\in H\) and \(z\in C\)

  1. (1)

    \(\langle x-P_C x, z-P_C x\rangle \le 0\);

  2. (2)

    \(P_C\) and \(I-P_C\) are both nonexpansive;

  3. (3)

    \(P_C\) and \(I-P_C\) are both firmly nonexpansive;

  4. (4)

    \(P_C\) and \(I-P_C\) are both 1-inverse strongly monotone.

We shall make use of the following well-known results.

Lemma 2.3

[25] Let \(\{x_n\}\) and \(\{z_n\}\) be bounded sequences in a Banach space X and let \(\{\gamma _n\}\) be a sequence in [0, 1] which satisfies the following condition:

$$\begin{aligned}0<\liminf _{n\rightarrow \infty }\gamma _n\le \limsup _{n\rightarrow \infty }\gamma _n<1.\end{aligned}$$

Suppose

$$\begin{aligned}x_{n+1}=\gamma _nx_n+(1-\gamma _n)z_n, \quad n\ge 0,\end{aligned}$$

and

$$\begin{aligned}\limsup _{n\rightarrow \infty }(\Vert z_{n+1}-z_n\Vert -\Vert x_{n+1}-x_n\Vert )\le 0.\end{aligned}$$

Then, \(\lim _{n\rightarrow \infty }\Vert z_n-x_n\Vert =0\).

Lemma 2.4

[2] Let H be a Hilbert space, C a closed convex subset of H, and \(T:C\rightarrow C\) a nonexpansive mapping with \(F(T)\ne \emptyset \). If \(\{x_n\}\) is a sequence in C weakly converging to x and if \(\{(I-T)x_n\}\) converges strongly to y, then \((I-T)x=y\).

We need the following lemma whose proof is elementary and therefore omitted.

Lemma 2.5

Let H be a Hilbert space. Then

  1. (i)

    \(\Vert x+y\Vert ^2\le \Vert x\Vert ^2+2\langle y, x+y\rangle , \quad \forall x, y\in H.\)

  2. (ii)

    \(\Vert tx+sy\Vert ^2=t(t+s)\Vert x\Vert ^2+s(t+s)\Vert y\Vert ^2-st\Vert x-y\Vert ^2, \quad \forall x, y\in H, \forall s, t\in {\mathbb {R}}\).

The following result is also well known. It plays a central role in our strong convergence result.

Lemma 2.6

[30] Let \(\{a_n\}\) be a sequence of nonnegative real numbers satisfying the following relation:

$$\begin{aligned}a_{n+1}\le (1-\alpha _n)a_n+\alpha _n\sigma _n+\gamma _n, \ n\ge 1,\end{aligned}$$

where (a) \(\{a_n\}\subset [0,1]\), \(\sum _{n=1}^\infty \alpha _n=\infty \); (b) \(\limsup \sigma _n\le 0\); and (c) \(\gamma _n \ge 0 (n\ge 1)\), \(\sum _{n=1}^\infty \gamma _n<\infty \). Then, \(a_n\rightarrow 0\) as \(n\rightarrow \infty \).

3 The generalized Halpern iteration with new control conditions

In this section, we investigate the strong convergence of the generalized Halpern iteration (1.6) with new control conditions on parameters. First, we give the equivalent characterization of condition (C4). We give the following two conditions (C5) and (C6), and prove that condition (C4), condition (C5), and condition (C6) are equivalent:

  1. (C5)

    \(0<\liminf _{n\rightarrow \infty }\alpha _n\le \limsup _{n\rightarrow \infty }\alpha _n<1\);

  2. (C6)

    \(0<\liminf _{n\rightarrow \infty }\beta _n\le \limsup _{n\rightarrow \infty }\beta _n<1\).

Lemma 3.1

Let \(\{\delta _n\}\), \(\{\alpha _n\}\) and \(\{\beta _n\}\) be real sequences in [0, 1], such that \(\delta _n+\alpha _n+\beta _n\le 1\). If condition (C1) and \(\lim _{n\rightarrow \infty }(1-\delta _n-\alpha _n-\beta _n)=0\) hold, then (C4) \(\Longleftrightarrow \) (C5) \(\Longleftrightarrow \) (C6).

Proof

We first prove that (C4) \(\Longleftrightarrow \) (C5).

\(\Longrightarrow \)’. From the monotonicity of the lower limit, we have

$$\begin{aligned}\liminf _{n\rightarrow \infty }\alpha _n\ge \liminf _{n\rightarrow \infty }\alpha _n\beta _n>0 \quad \text {and} \quad \liminf _{n\rightarrow \infty }\beta _n\ge \liminf _{n\rightarrow \infty }\alpha _n\beta _n>0.\end{aligned}$$

Since \(\alpha _n\le 1-\delta _n-\beta _n\), we obtain that

$$\begin{aligned} \limsup _{n\rightarrow \infty }\alpha _n&\le \limsup _{n\rightarrow \infty }(1-\delta _n-\beta _n)\\&=\limsup _{n\rightarrow \infty }(1-\delta _n)-\liminf _{n\rightarrow \infty }\beta _n\\&=1-\liminf _{n\rightarrow \infty }\beta _n<1. \end{aligned}$$

\(\Longleftarrow \)’. If \(0<\liminf _{n\rightarrow \infty }\alpha _n\le \limsup _{n\rightarrow \infty }\alpha _n<1\), then

$$\begin{aligned} \liminf _{n\rightarrow \infty }\beta _n&= \liminf _{n\rightarrow \infty }(1-\delta _n-(1-\delta _n-\alpha _n-\beta _n)-\alpha _n)\\&=\liminf _{n\rightarrow \infty }(1-\delta _n-(1-\delta _n-\alpha _n-\beta _n))-\limsup _{n\rightarrow \infty }\alpha _n\\&=1-\limsup _{n\rightarrow \infty }\alpha _n>0. \end{aligned}$$

Therefore, \(\liminf _{n\rightarrow \infty }\alpha _n\beta _n\ge \liminf _{n\rightarrow \infty }\alpha _n\cdot \liminf _{n\rightarrow \infty }\beta _n>0\).

In the same way, we can prove that (C4) \(\Longleftrightarrow \) (C6). \(\square \)

Second, we give the following two conditions (C7) and (C8). By replacing conditions (C2) and (C3) with conditions (C7) and (C8), respectively, we prove the strong convergence of the generalized Halpern iteration (1.6)

  1. (C7)

    \(1-\delta _n-\alpha _n-\beta _n=o(\delta _n), \ \mathrm {i.e.},\ \lim _{n\rightarrow \infty }\frac{1-\delta _n-\alpha _n-\beta _n}{\delta _n}=0\);

  2. (C8)

    \(\Vert r_n\Vert =o(\delta _n),\ \mathrm {i.e.},\ \lim _{n\rightarrow \infty }\frac{\Vert r_n\Vert }{\delta _n}=0\).

Theorem 3.2

Let H be a real Hilbert space, K a nonempty, closed, and convex subset of H, \(T:H\rightarrow K\) a nonexpansive mapping with \(F(T)\ne \emptyset \). Let \(\{x_n\}\) be generated by the iteration (1.6). Assume that \(\{\delta _n\}\), \(\{\alpha _n\}\) and \(\{\beta _n\}\) are real sequences in [0, 1], such that \(\delta _n+\alpha _n+\beta _n\le 1\) for all \(n\ge 0\) and conditions (C1),(C4),(C7), and (C8) hold. Then, \(\{x_n\}\) strongly converges to z, where \(z:=P_{F(T)}u\).

Proof

It is easy to know that the projection \(z=P_{F(T)}u\) exists, since F(T) is nonempty, closed, and convex. We now divide the proof into five steps.

Step 1. We show that \(\{x_n\}\) is bounded. In fact, without losing generality, it can be assumed from condition (C8) that \(\Vert r_n\Vert \le \delta _n \ (n\ge 0)\). Take \(x^*\in F(T)\). It follows from (1.6) that:

$$\begin{aligned}&\Vert x_{n+1}-x^*\Vert \\ {}&\quad =\Vert \delta _n(u-x^*)+\alpha _n(x_n-x^*)+\beta _n(Tx_n-x^*)+r_n-(1-\delta _n-\alpha _n-\beta _n)x^*\Vert \\&\quad \le \delta _n\Vert u-x^*\Vert +\alpha _n\Vert x_n\!-\!x^*\Vert \!+\!\beta _n\Vert Tx_n\!-\!x^*\Vert +\Vert r_n\Vert \!+\!(1-\delta _n-\alpha _n-\beta _n)\Vert x^*\Vert \\&\quad \le \delta _n\Vert u-x^*\Vert +(\alpha _n+\beta _n)\Vert x_n-x^*\Vert +\delta _n+(1-\delta _n-\alpha _n-\beta _n)\Vert x^*\Vert \\&\quad = \delta _n(\Vert u-x^*\Vert +1)+(\alpha _n+\beta _n)\Vert x_n-x^*\Vert +(1-\delta _n-\alpha _n-\beta _n)\Vert x^*\Vert \\&\quad \le \max \{\Vert u-x^*\Vert +1, \Vert x_n-x^*\Vert , \Vert x^*\Vert \}\\&\quad \le \max \{\Vert u\Vert +\Vert x^*\Vert +1, \Vert x_n-x^*\Vert \}. \end{aligned}$$

Using induction, it is easy to see that this implies

$$\begin{aligned}\Vert x_{n+1}-x^*\Vert \le \max \{\Vert u\Vert +\Vert x^*\Vert +1, \Vert x_0-x^*\Vert \}.\end{aligned}$$

Therefore, \(\{x_n\}\) is bounded in H, so is \(\{Tx_n\}\).

Step 2. Here, we show that \(\Vert x_{n+1}-x_n\Vert \rightarrow 0\). Since condition (C4) holds, it can be obtained from Lemma 3.1 that

$$\begin{aligned}0<\liminf _{n\rightarrow \infty }\alpha _n\le \limsup _{n\rightarrow \infty }\alpha _n<1.\end{aligned}$$

Setting \(z_{n}=\frac{\delta _n u+\beta _n Tx_n+r_n}{1-\alpha _n}\), we have

$$\begin{aligned} x_{n+1}=\alpha _n x_n+(1-\alpha _n)z_n. \end{aligned}$$
(3.1)

Observe that

$$\begin{aligned} z_{n+1}-z_n&=\frac{\delta _{n+1} u+\beta _{n+1} Tx_{n+1}+r_{n+1}}{1-\alpha _{n+1}}-\frac{\delta _n u+\beta _n Tx_n+r_n}{1-\alpha _n}\\&=\Big (\frac{\delta _{n+1} u+r_{n+1}}{1-\alpha _{n+1}}-\frac{\delta _n u+r_n}{1-\alpha _n}\Big )+(Tx_{n+1}-Tx_n)\\&\quad +\Big (\frac{\beta _{n+1} Tx_{n+1}}{1-\alpha _{n+1}}-Tx_{n+1}\Big )-\Big (\frac{\beta _n Tx_n}{1-\alpha _n}-Tx_n\Big )\\&=\Big (\frac{\delta _{n+1} u+r_{n+1}}{1-\alpha _{n+1}}-\frac{\delta _n u+r_n}{1-\alpha _n}\Big )+(Tx_{n+1}-Tx_n)\\&\quad -\frac{(1-\alpha _{n+1}-\beta _{n+1}) Tx_{n+1}}{1-\alpha _{n+1}}+\frac{(1-\alpha _n-\beta _n) Tx_n}{1-\alpha _n}. \end{aligned}$$

This implies that

$$\begin{aligned}&\Vert z_{n+1}-z_n\Vert -\Vert x_{n+1}-x_n\Vert \\&\quad \le \frac{\delta _{n+1} \Vert u\Vert +\Vert r_{n+1}\Vert }{1-\alpha _{n+1}}+\frac{\delta _n \Vert u\Vert +\Vert r_n\Vert }{1-\alpha _n}+\Vert Tx_{n+1}-Tx_n\Vert \\&\qquad +\frac{(1-\alpha _{n+1}-\beta _{n+1}) \Vert Tx_{n+1}\Vert }{1-\alpha _{n+1}}+\frac{(1-\alpha _n-\beta _n) \Vert Tx_n\Vert }{1-\alpha _n}\\&\qquad -\Vert x_{n+1}-x_n\Vert \\&\quad \le \frac{\delta _{n+1} \Vert u\Vert +\Vert r_{n+1}\Vert }{1-\alpha _{n+1}}+\frac{\delta _n \Vert u\Vert +\Vert r_n\Vert }{1-\alpha _n}\\&\qquad +\frac{\delta _{n+1}+(1-\delta _{n+1}-\alpha _{n+1}-\beta _{n+1}) }{1-\alpha _{n+1}}\Vert Tx_{n+1}\Vert \\&\qquad +\frac{\delta _n+(1-\delta _n-\alpha _n-\beta _n) }{1-\alpha _n}\Vert Tx_n\Vert . \end{aligned}$$

Since \(\limsup _{n\rightarrow \infty }\alpha _n<1, \lim _{n\rightarrow \infty }\delta _n=0, \lim _{n\rightarrow \infty }\Vert r_n\Vert =0, \lim _{n\rightarrow \infty }(1-\delta _n-\alpha _n-\beta _n)=0\) and \(\{Tx_n\}\) is bounded, we obtain that

$$\begin{aligned}\limsup _{n\rightarrow \infty }(\Vert z_{n+1}-z_n\Vert -\Vert x_{n+1}-x_n\Vert )\le 0.\end{aligned}$$

Hence, by Lemma 2.3, we have

$$\begin{aligned}\lim _{n\rightarrow \infty }\Vert z_n-x_n\Vert =0.\end{aligned}$$

It follows from (3.1) that:

$$\begin{aligned} \lim _{n\rightarrow \infty }\Vert x_{n+1}-x_n\Vert =0. \end{aligned}$$
(3.2)

Step 3. We now show that \(\Vert x_n-Tx_n\Vert \rightarrow 0\). Indeed, we have

$$\begin{aligned} \Vert x_n-Tx_n\Vert&\le \Vert x_{n+1}-Tx_n\Vert +\Vert x_{n+1}-x_n\Vert \\&=\Vert \delta _n(u-Tx_n)+\alpha _n(x_n-Tx_n)-(1-\delta _n-\alpha _n-\beta _n)Tx_n+r_n\Vert \\&\quad +\Vert x_{n+1}-x_n\Vert \\&\le \delta _n\Vert u-Tx_n\Vert +\alpha _n\Vert x_n-Tx_n\Vert +(1-\delta _n-\alpha _n-\beta _n)\Vert Tx_n\Vert \\&\quad +\Vert r_n\Vert +\Vert x_{n+1}-x_n\Vert . \end{aligned}$$

Hence

$$\begin{aligned} \Vert x_n-Tx_n\Vert&\le \frac{1}{1-\alpha _n}\big (\Vert x_{n+1}-x_n\Vert +\delta _n\Vert u-Tx_n\Vert \\&\quad +(1-\delta _n-\alpha _n-\beta _n)\Vert Tx_n\Vert +\Vert r_n\Vert \big ). \end{aligned}$$

From (3.2) and conditions (C1), (C7), and (C8), this implies that

$$\begin{aligned} \Vert x_n-Tx_n\Vert \rightarrow 0. \end{aligned}$$
(3.3)

Step 4. We show that \(\limsup _{n\rightarrow \infty }\langle u-z, x_{n+1}-z\rangle \le 0\). Since \(\{x_n\}\) is bounded, we can extract a subsequence \(\{x_{n_k}\}\) of \(\{x_n\}\), such that

$$\begin{aligned}\limsup _{n\rightarrow \infty }\langle u-z, x_n-z\rangle =\lim _{k\rightarrow \infty }\langle u-z, x_{n_k}-z\rangle ,\end{aligned}$$

and \(x_{n_k}\rightharpoonup {\bar{x}}\). It follows from Lemma 2.4 and (3.3) that \({\bar{x}}\in F(T)\). Therefore, we obtain from Lemma 2.2 (i) that

$$\begin{aligned} \limsup _{n\rightarrow \infty }\langle u-z, x_{n+1}-z\rangle&=\limsup _{n\rightarrow \infty }\langle u-z, x_{n}-z\rangle \\&=\lim _{k\rightarrow \infty }\langle u-z, x_{n_k}-z\rangle \\&=\langle u-z, {\bar{x}}-z\rangle \le 0. \end{aligned}$$

Step 5. In this final step, we prove that \(x_n\rightarrow z\). From Lemma 2.5 (i) and (ii), we have

$$\begin{aligned} \Vert x_{n+1}-z\Vert ^2&\le \Vert \alpha _n(x_n-z)+\beta _n(Tx_n-z)\Vert ^2\nonumber \\&\quad +2\langle \delta _n(u-z)+r_n-(1-\delta _n-\alpha _n-\beta _n)z, x_{n+1}-z\rangle \nonumber \\&= \alpha _n(\alpha _n+\beta _n)\Vert x_n-z\Vert ^2+\beta _n(\alpha _n+\beta _n)\Vert Tx_n-z\Vert ^2\nonumber \\&\quad -\alpha _n\beta _n\Vert x_n-Tx_n\Vert ^2+2\delta _n\langle u-z, x_{n+1}-z\rangle \nonumber \\&\quad +2\langle r_n, x_{n+1}-z\rangle -2(1-\delta _n-\alpha _n-\beta _n)\langle z, x_{n+1}-z\rangle \nonumber \\&\le (\alpha _n+\beta _n)^2\Vert x_n-z\Vert ^2+2\delta _n\langle u-z, x_{n+1}-z\rangle \nonumber \\&\quad +2\langle r_n, x_{n+1}-z\rangle -2(1-\delta _n-\alpha _n-\beta _n)\langle z, x_{n+1}-z\rangle \nonumber \\&\le (1-\delta _n)\Vert x_n-z\Vert ^2+2\delta _n\langle u-z, x_{n+1}-z\rangle \nonumber \\&\quad +2\langle r_n, x_{n+1}-z\rangle -2(1-\delta _n-\alpha _n-\beta _n)\langle z, x_{n+1}-z\rangle \nonumber \\&\le (1-\delta _n)\Vert x_n-z\Vert ^2+2\delta _n\langle u-z, x_{n+1}-z\rangle \nonumber \\&\quad +\Vert r_n\Vert M_1 +(1-\delta _n-\alpha _n-\beta _n)M_1, \end{aligned}$$
(3.4)

for some \(M_1>0\) (recall that \(\{x_n\}\) is bounded). Applying Lemma 2.6 to (3.4) and using conditions (C1), (C7), and (C8), we have that \(\lim _{n\rightarrow \infty }\Vert x_n-z\Vert =0\). Thus, \(x_n\rightarrow z=P_{F(T)}u\) whenever \(n\rightarrow \infty \). \(\square \)

Remark 3.3

(1) Condition (C7) and condition (C2) are not comparable. Indeed, take \(\delta _n=\frac{1}{\sqrt{n}}\), \(\alpha _n=1-\frac{1}{\sqrt{n}}-\frac{2}{n}-\epsilon _0\), \(\beta _n=\frac{1}{n}+\epsilon _0\) and \(r_n\equiv 0\) for all \(n\ge 5\), where \(\epsilon _0\) is a constant satisfying \(0<\epsilon _0<\frac{3-\sqrt{5}}{5}\). Then, we have \(1-\delta _n-\alpha _n-\beta _n=\frac{1}{n} \ (n\ge 5)\). It is easy to verify that condition (C7) is satisfied, but condition (C2) is violated. On the other hand, we take

$$\begin{aligned}\delta _n=\left\{ \begin{array}{ll} \frac{1}{n^2}, &{}\quad \text {if} \ n \ \text {is odd}, \\ \frac{1}{n}, &{}\quad \text {if} \ n \ \text {is even}, \end{array} \right. \\\alpha _n=\left\{ \begin{array}{ll} 1-\frac{3}{n^2}-\epsilon _0, &{}\quad \text {if} \ n \ \text {is odd}, \\ 1-\frac{2}{n}-\epsilon _0, &{}\quad \text {if} \ n \ \text {is even}, \end{array} \right. \\\beta _n=\left\{ \begin{array}{ll} \frac{1}{n^2}+\epsilon _0, &{}\quad \text {if} \ n \ \text {is odd}, \\ \frac{1}{n}+\epsilon _0, &{}\quad \text {if} \ n \ \text {is even}, \end{array} \right. \end{aligned}$$

and \(r_n\equiv 0\) for all \(n\ge 3\), where \(\epsilon _0\) is a constant satisfying \(0<\epsilon _0<\frac{1}{2}\). Then

$$\begin{aligned}1-\delta _n-\alpha _n-\beta _n=\left\{ \begin{array}{ll} \frac{1}{n^2}, &{}\quad \text {if} \ n \ \text {is odd}, \\ 0, &{}\quad \text {if} \ n \ \text {is even}, \end{array} \right. \end{aligned}$$

which satisfies condition (C2), but does not satisfy (C7).

(2) We now show that condition (C8) and condition (C3) are also not comparable. Take \(\delta _n=\frac{1}{\sqrt{n}}\), \(\alpha _n=1-\frac{1}{\sqrt{n}}-\frac{1}{n}-\epsilon _0\), \(\beta _n=\frac{1}{n}+\epsilon _0\) and \(\Vert r_n\Vert =\frac{1}{n}\) for all \(n\ge 3\), where \(\epsilon _0\) is a constant satisfying \(0<\epsilon _0<\frac{2-\sqrt{3}}{3}\). It is easy to verify that condition (C8) is satisfied, but condition (C3) is violated. On the other hand, we take

$$\begin{aligned}\delta _n=\left\{ \begin{array}{ll} \frac{1}{n^2}, &{}\quad \text {if} \ n \ \text {is odd}, \\ \frac{1}{n}, &{}\quad \text {if} \ n \ \text {is even}, \end{array} \right. \\\alpha _n=\left\{ \begin{array}{ll} 1-\frac{2}{n^2}-\epsilon _0, &{}\quad \text {if} \ n \ \text {is odd}, \\ 1-\frac{2}{n}-\epsilon _0, &{}\quad \text {if} \ n \ \text {is even}, \end{array} \right. \\\beta _n=\left\{ \begin{array}{ll} \frac{1}{n^2}+\epsilon _0, &{}\quad \text {if} \ n \ \text {is odd}, \\ \frac{1}{n}+\epsilon _0, &{}\quad \text {if} \ n \ \text {is even}, \end{array} \right. \end{aligned}$$

and \(\Vert r_n\Vert =\frac{1}{n^2}\) for all \(n\ge 3\), where \(\epsilon _0\) is a constant satisfying \(0<\epsilon _0<\frac{1}{2}\). It is easy to verify that condition (C3) is satisfied, but condition (C8) is violated.

(3) Conditions (C7) and (C8) broaden the application range of the generalized Halpern iteration. In practical applications, when the residual \(\Vert r_n\Vert \) does not satisfy condition (C3), the strong convergence of the generalized Halpern iteration can still be guaranteed by adjusting the parameter \(\delta _n\) to satisfy condition (C8).

It is easy to observe that condition (C4) is very harsh, which excludes many common choices, such as \(\delta _n=\frac{1}{2(n+1)}, \alpha _n=\frac{1}{2(n+1)}, \beta _n=1-\frac{1}{n+1}\) and \(r_n\equiv 0\). Therefore, we give the following new control condition, that includes such a choice as a special case:

  1. (C9)

    \(\limsup _{n\rightarrow \infty }\alpha _n<1\), either \(\lim _{n\rightarrow \infty }\frac{\delta _{n-1}}{\delta _n}=1\) or \(\sum _{n=1}^\infty |\delta _n-\delta _{n-1}|<\infty \), either \(\lim _{n\rightarrow \infty }\frac{|\alpha _n-\alpha _{n-1}|}{\delta _n}=0\) or \(\sum _{n=1}^\infty |\alpha _n-\alpha _{n-1}|<\infty \).

Theorem 3.4

Let H be a real Hilbert space, K a nonempty, closed, and convex subset of H, and \(T:H\rightarrow K\) a nonexpansive mapping with \(F(T)\ne \emptyset \). Let \(\{x_n\}\) be generated by the iteration (1.6). Assume that \(\{\delta _n\}\), \(\{\alpha _n\}\) and \(\{\beta _n\}\) are real sequences in [0, 1], such that \(\delta _n+\alpha _n+\beta _n\le 1\) for all \(n\ge 0\) and conditions (C1–C3) and (C9) hold. Then, \(\{x_n\}\) strongly converges to z, where \(z:=P_{F(T)}u\).

Proof

Similar to the proof of Theorem 3.2, we also divide the proof into five steps.

Step 1. We show that \(\{x_n\}\) is bounded. Take \(x^*\in F(T)\). It follows from (1.6) that:

$$\begin{aligned}&\Vert x_{n+1}-x^*\Vert \\&\quad =\Vert \delta _n(u-x^*)+\alpha _n(x_n-x^*)+\beta _n(Tx_n-x^*)+r_n-(1-\delta _n-\alpha _n-\beta _n)x^*\Vert \\&\quad \le \delta _n\Vert u-x^*\Vert +\alpha _n\Vert x_n-x^*\Vert +\beta _n\Vert Tx_n-x^*\Vert +\Vert r_n\Vert +(1-\delta _n-\alpha _n-\beta _n)\Vert x^*\Vert \\&\quad \le \delta _n\Vert u-x^*\Vert +(\alpha _n+\beta _n)\Vert x_n-x^*\Vert +(1-\delta _n-\alpha _n-\beta _n)\Vert x^*\Vert +\Vert r_n\Vert \\&\quad \le \max \{\Vert u-x^*\Vert , \Vert x_n-x^*\Vert , \Vert x^*\Vert \}+\Vert r_n\Vert \\&\quad \le \max \{\Vert u\Vert +\Vert x^*\Vert , \Vert x_n-x^*\Vert \}+\Vert r_n\Vert . \end{aligned}$$

Using induction, it is easy to see that this implies

$$\begin{aligned}\Vert x_{n+1}-x^*\Vert \le \max \{\Vert u\Vert +\Vert x^*\Vert , \Vert x_0-x^*\Vert \}+\sum _{k=0}^n\Vert r_k\Vert ,\end{aligned}$$

which yields

$$\begin{aligned}\Vert x_{n+1}-x^*\Vert \le \max \{\Vert u\Vert +\Vert x^*\Vert , \Vert x_0-x^*\Vert \}+\sum _{k=0}^{\infty }\Vert r_k\Vert .\end{aligned}$$

It follows from condition (C3) that \(\{x_n\}\) is bounded in H, so is \(\{Tx_n\}\).

Step 2. We now show that \(\Vert x_{n+1}-x_n\Vert \rightarrow 0\). Indeed, from (1.6), we have

$$\begin{aligned} \Vert x_{n+1}-x_n\Vert&=\Vert \alpha _n(x_n-x_{n-1})+\beta _n(Tx_n-Tx_{n-1})+(\delta _n-\delta _{n-1})(u-Tx_{n-1})\\&\quad +(\alpha _n-\alpha _{n-1})(x_{n-1}-Tx_{n-1})+(r_n-r_{n-1})\\&\quad +\big [(1-\delta _{n-1}-\alpha _{n-1}-\beta _{n-1})-(1-\delta _n-\alpha _n-\beta _n)\big ]Tx_{n-1}\Vert \\&\le \alpha _n\Vert x_n-x_{n-1}\Vert +\beta _n\Vert Tx_n-Tx_{n-1}\Vert +|\delta _n-\delta _{n-1}| \ \Vert u-Tx_{n-1}\Vert \\&\quad +|\alpha _n-\alpha _{n-1}| \ \Vert x_{n-1}-Tx_{n-1}\Vert +\Vert r_n-r_{n-1}\Vert \\&\quad +|(1-\delta _{n-1}-\alpha _{n-1}-\beta _{n-1})-(1-\delta _n-\alpha _n-\beta _n)| \ \Vert Tx_{n-1}\Vert \\&\le (\alpha _n+\beta _n)\Vert x_n-x_{n-1}\Vert +|\delta _n-\delta _{n-1}| \ \Vert u-Tx_{n-1}\Vert \\&\quad +|\alpha _n-\alpha _{n-1}| \ \Vert x_{n-1}-Tx_{n-1}\Vert +\Vert r_n-r_{n-1}\Vert \\&\quad +|(1-\delta _{n-1}-\alpha _{n-1}-\beta _{n-1})-(1-\delta _n-\alpha _n-\beta _n)| \ \Vert Tx_{n-1}\Vert \\&\le (1-\delta _n)\Vert x_n-x_{n-1}\Vert +\big (|\delta _n-\delta _{n-1}|+|\alpha _n-\alpha _{n-1}|\big ) \ M_2\\&\qquad +\Vert r_n-r_{n-1}\Vert \\&\quad +|(1-\delta _{n-1}-\alpha _{n-1}-\beta _{n-1})-(1-\delta _n-\alpha _n-\beta _n)| \ M_2, \end{aligned}$$

for some \(M_2>0\) (recall that \(\{x_n\}\) and \(\{Tx_n\}\) are bounded). Since \(\sum _{n=0}^\infty \Vert r_n\Vert <\infty \) and \(\sum _{n=0}^\infty (1-\delta _n-\alpha _n-\beta _n)<\infty \), it is easy to obtain that \(\sum _{n=1}^\infty \Vert r_n-r_{n-1}\Vert <\infty \) and \(\sum _{n=1}^\infty |(1-\delta _{n-1}-\alpha _{n-1}-\beta _{n-1})-(1-\delta _n-\alpha _n-\beta _n)|<\infty \). From conditions (C1), (C9) and Lemma 2.6, we have

$$\begin{aligned}\Vert x_{n+1}-x_n\Vert \rightarrow 0.\end{aligned}$$

The proof of Step 3–Step 5 is the same as that of Theorem 3.2, so we omit it. \(\square \)

The following example shows that our Theorem 3.4 fails if condition (C9) does not hold.

Example 3.5

Let H be any real Hilbert space. Choose an arbitrary element \(y\in H\), such that \(y\ne 0\) and \(\Vert y\Vert =1\), and define a subset K of H by \(K=\{x\in H\mid x=\lambda y, \lambda \in [0,1]\}\), i.e., K is the connecting line from the origin to the given point y. Observe that K is a nonempty, closed, and convex subset of H. Then, define the mapping \(T:H\rightarrow K\) by \(Tx=0\) for all \(x\in H\). Clearly, T is nonexpansive, and \(F(T)=\{0\}\).

(1) Consider the iterative scheme (1.6) with

$$\begin{aligned}\delta _n=\frac{1}{2(n+1)}, \alpha _n=1-\frac{1}{n+1}, \beta _n=\frac{1}{2(n+1)}, r_n\equiv 0, \forall n\ge 0.\end{aligned}$$

Take \(u=y\in K\), and let \(x_0=u\) be the starting point. It is clear that conditions (C1–C3) are satisfied, but both condition (C9) and condition (C4) are violated. It is proved that \(\{x_n\}\) generated by the iteration (1.6) does not strongly converge to 0, which is the unique fixed point of T in [12].

(2) Consider the iterative scheme (1.6) with

$$\begin{aligned}\delta _n=\frac{1}{2(n+1)}, \alpha _n=\frac{1}{2(n+1)}, \beta _n=1-\frac{1}{n+1}, r_n\equiv 0, \forall n\ge 0.\end{aligned}$$

Then, it is clear that conditions (C1–C3) are satisfied, but condition (C4) is violated. On the other hand, since condition (C9) is satisfied, we verifies that \(\{x_n\}\) generated by the iteration (1.6) strongly converges to 0, which is the unique fixed point of T for all \(u\in K\) and \(x_0\in H\). It follows from (1.6) that:

$$\begin{aligned} \Vert x_{n+1}\Vert&=\Vert \frac{1}{2(n+1)}u+\frac{1}{2(n+1)}x_n\Vert \\&\le \frac{1}{2(n+1)}\Vert u\Vert +\frac{1}{2(n+1)}\Vert x_n\Vert \\&\le \frac{1}{2(n+1)}\Vert u\Vert +\left( 1-\frac{1}{2(n+1)}\right) \Vert x_n\Vert \\&\le \max \{\Vert u\Vert , \Vert x_n\Vert \}. \end{aligned}$$

Using induction, it is easy to obtain

$$\begin{aligned}\Vert x_{n+1}\Vert \le \max \{\Vert u\Vert , \Vert x_0\Vert \}.\end{aligned}$$

Thus, \(\{x_n\}\) is bounded. This implies that

$$\begin{aligned}\Vert x_{n+1}\Vert \le \frac{1}{2(n+1)}\Vert u\Vert +\frac{1}{2(n+1)}\Vert x_n\Vert \rightarrow 0, \quad (n\rightarrow \infty ).\end{aligned}$$

Hence, \(\{x_n\}\) strongly converges to 0.

Remark 3.6

(1) It is easy to verify that the canonical choices \(\delta _n=\frac{1}{2(n+1)}, \alpha _n=\frac{1}{2(n+1)}, \beta _n=1-\frac{1}{n+1}\) and \(r_n\equiv 0\) satisfy the conditions (C1–C3) and (C9), but do not satisfy the condition (C4).

(2) In conditions (C1–C3) and (C9), if \(\delta _n+\alpha _n+\beta _n=1\), \(\alpha _n\equiv 0\) and \(r_n\equiv 0\), then conditions (C1–C3) and (C9) are reduced to conditions (A1), (A3) or (A4). Therefore, iteration (1.6) with conditions (C1–C3) and (C9) is a generalization of iteration (1.2) with conditions (A1), (A3), or (A4). However, \(\alpha _n\equiv 0\) does not satisfy condition (C4). Therefore, iteration (1.6) with conditions (C1–C4) is not a generalization of iteration (1.2).

(3) We show that condition (C9) and condition (C4) are not comparable. We take \(\delta _n=\frac{1}{n+3}\), \(\alpha _n=\alpha _n^\prime +\epsilon _0\), \(\beta _n=1-\frac{1}{n+3}-\alpha _n^\prime -\epsilon _0\) and \(r_n\equiv 0\) for all \(n\ge 0\), where

$$\begin{aligned} \alpha _n^\prime =\left\{ \begin{array}{ll} \frac{1}{n+2}, &{}\quad \text {if} \ n \ \text {is even}, \\ \frac{1}{\sqrt{n+3}}, &{}\quad \text {if} \ n \ \text {is odd}, \end{array} \right. \end{aligned}$$

\(\epsilon _0\) is a constant satisfying \(0<\epsilon _0<\frac{1}{6}\). It is easy to verify that conditions (C1–C4) are satisfied. However, condition (C9) is violated. Indeed, we just need to verify that neither \(\lim _{n\rightarrow \infty }\frac{|\alpha _n-\alpha _{n-1}|}{\delta _n}=0\) nor \(\sum _{n=0}^\infty |\alpha _n-\alpha _{n-1}|<\infty \) holds. It is easy to obtain that

$$\begin{aligned} |\alpha _n-\alpha _{n-1}|=\left\{ \begin{array}{ll} \frac{1}{\sqrt{n+2}}-\frac{1}{n+2}, &{}\quad \text {if} \ n \ \text {is even}, \\ \frac{1}{\sqrt{n+3}}-\frac{1}{n+1}, &{}\quad \text {if} \ n \ \text {is odd}. \end{array} \right. \end{aligned}$$

Thus

$$\begin{aligned} \frac{|\alpha _n-\alpha _{n-1}|}{\delta _n}=\left\{ \begin{array}{ll} \frac{(n+2-\sqrt{n+2})(n+3)}{\sqrt{n+2}(n+2)}, &{}\quad \text {if} \ n \ \text {is even}, \\ \frac{(n+1-\sqrt{n+3})(n+3)}{(n+1)\sqrt{n+3}}, &{}\quad \text {if} \ n \ \text {is odd}. \end{array} \right. \end{aligned}$$

This implies that \(\frac{|\alpha _n-\alpha _{n-1}|}{\delta _n}\rightarrow \infty \). When \(n>0\) and n is even

$$\begin{aligned} |\alpha _n-\alpha _{n-1}|= \frac{1}{\sqrt{n+2}}-\frac{1}{n+2}>\frac{1}{\sqrt{n+3}}-\frac{1}{n+1}. \end{aligned}$$

Hence

$$\begin{aligned}\sum _{n=0}^\infty |\alpha _n-\alpha _{n-1}|>\big (\frac{1}{\sqrt{2}}-\frac{1}{2}\big )+ \sum _{n=1}^\infty \big (\frac{1}{\sqrt{n+3}}-\frac{1}{n+1}\big ).\end{aligned}$$

This implies that \(\sum _{n=0}^\infty |\alpha _n-\alpha _{n-1}|\) is divergent, since \(\sum _{n=1}^\infty \big (\frac{1}{\sqrt{n+3}}-\frac{1}{n+1}\big )\) is divergent.

From (1) of Remark 3.6, it can be seen that condition (C9) and condition (C4) are not comparable.

Next, we give another control condition on parameters. Under this condition, the strong convergence of \(\{x_n\}\) generated by the iteration (1.6) to a fixed point of T can also be proved.

  1. (C10)

    \(\liminf _{n\rightarrow \infty }\beta _n>0\), either \(\lim _{n\rightarrow \infty }\frac{\delta _{n-1}}{\delta _n}=1\) or \(\sum _{n=1}^\infty |\delta _n-\delta _{n-1}|<\infty \), either \(\lim _{n\rightarrow \infty }\frac{|\beta _n-\beta _{n-1}|}{\delta _n}=0\) or \(\sum _{n=1}^\infty |\beta _n-\beta _{n-1}|<\infty \).

Theorem 3.7

Let H be a real Hilbert space, K a nonempty, closed, and convex subset of H, and \(T:H\rightarrow K\) a nonexpansive mapping with \(F(T)\ne \emptyset \). Let \(\{x_n\}\) be generated by the iteration (1.6). Assume that \(\{\delta _n\}\), \(\{\alpha _n\}\) and \(\{\beta _n\}\) are real sequences in [0, 1], such that \(\delta _n+\alpha _n+\beta _n\le 1\) for all \(n\ge 0\) and conditions (C1)-(C3) and (C10) hold. Then, \(\{x_n\}\) strongly converges to z, where \(z:=P_{F(T)}u\).

The proof of Theorem 3.7 is similar to that of Theorem 3.4, so we omit its proof.

4 Convergence rate analysis for the generalized Halpern iteration

In [8], Cominetti et al. showed that \(\Vert x_n-Tx_n\Vert \) in (1.1) converges to zero at a rate of \(O(1/\sqrt{\sigma _n})\)(big-O), where \(\sigma _n=\sum _{k=0}^{n}\lambda _k(1-\lambda _k)\), \(n\ge 0\), whenever \(\lim _{n\rightarrow \infty }\sigma _n=\infty \). In particular, when \(\lambda _n=\lambda \) (\(\lambda \in (0,1)\)), this convergence rate becomes \(O(1/\sqrt{n+1})\). In this section, we will give the convergence rate of \(\Vert x_n-Tx_n\Vert \) and \(\Vert x_n-x_{n-1}\Vert \) in (1.6) with a particular choice of parameters \(\delta _n, \alpha _n, \beta _n\) and \(r_n\). The technical lemma which we state next, and for which we refer to [22], will play a crucial role in the convergence rate analysis.

Lemma 4.1

Let \(M>0\). Assume that \(\{a_n\}\) is a sequence of nonnegative real numbers which satisfy \(a_1\le 2M\) and

$$\begin{aligned}a_{n+1}\le (1-b_n)a_n +(b_{n-1}-b_n)\cdot 4M, \quad n\ge 1,\end{aligned}$$

where \(\{b_n\}\) is a sequence which is defined by \(b_n=\frac{2}{n+4}\). Then, the sequence \(\{a_n\}\) satisfies

$$\begin{aligned}a_n\le \frac{8M}{n+3}, \quad n\ge 1.\end{aligned}$$

Proof

We will assume that the claim is true for \(n\ge 1\) and prove that it is true for \(n+1\). Using the induction assumption, we obtain

$$\begin{aligned} a_{n+1}&\le (1-b_n)a_n +(b_{n-1}-b_n)\cdot 4M\\&\le \Big (1-\frac{2}{n+4}\Big )\frac{8M}{n+3}+\Big (\frac{2}{n+3}-\frac{2}{n+4}\Big )\cdot 4M\\&=\frac{8(n+2)M}{(n+3)(n+4)}+\frac{8M}{(n+3)(n+4)}\\&=\frac{8M}{n+4}. \end{aligned}$$

This completes the proof of the desired result. \(\square \)

Theorem 4.2

Let \(\{x_n\}\) be a sequence generated by the iteration (1.6), where \(\{\delta _n\}, \{\alpha _n\}, \{\beta _n\}\), and \(\{r_n\}\) are defined by

$$\begin{aligned}\delta _n=\alpha _n=\frac{2}{n+4}, \quad \beta _n=\frac{n}{n+4} \quad \text {and} \quad r_n\equiv 0, \quad n\ge 0.\end{aligned}$$

Then, for any \({\bar{x}}\in Fix(T)\), the two sequences \(\{\Vert x_n-x_{n-1}\Vert \}\) and \(\{\Vert x_n-Tx_n\Vert \}\) converge to 0, and the convergence rates are given by

$$\begin{aligned}\Vert x_n-x_{n-1}\Vert \le \frac{8C_{{\bar{x}}}}{n+3},\end{aligned}$$

and

$$\begin{aligned}\Vert x_n-Tx_n\Vert \le \frac{12C_{{\bar{x}}}}{n+2},\end{aligned}$$

where \(C_{{\bar{x}}}=\max \{\Vert u-{\bar{x}}\Vert , \Vert x_0-{\bar{x}}\Vert \}\).

Proof

Since \(\delta _n+\alpha _n+\beta _n=1\) and \(r_n\equiv 0\), it follows from (1.6) that:

$$\begin{aligned} \Vert x_{n+1}-{\bar{x}}\Vert&=\Vert \delta _n(u-{\bar{x}})+\alpha _n(x_n-{\bar{x}})+\beta _n(Tx_n-{\bar{x}})\Vert \\&\le \delta _n\Vert u-{\bar{x}}\Vert +\alpha _n\Vert x_n-{\bar{x}}\Vert +\beta _n\Vert Tx_n-{\bar{x}}\Vert \\&\le \delta _n\Vert u-{\bar{x}}\Vert +(\alpha _n+\beta _n)\Vert x_n-{\bar{x}}\Vert \\&= \delta _n\Vert u-{\bar{x}}\Vert +(1-\delta _n)\Vert x_n-{\bar{x}}\Vert \\&\le \max \{\Vert u-{\bar{x}}\Vert , \Vert x_n-{\bar{x}}\Vert \}. \end{aligned}$$

Using induction, we have

$$\begin{aligned}\Vert x_{n+1}-{\bar{x}}\Vert \le \max \{\Vert u-{\bar{x}}\Vert , \Vert x_0-{\bar{x}}\Vert \}=C_{{\bar{x}}}.\end{aligned}$$

Thus, we obtain that

$$\begin{aligned} \Vert u-Tx_{n-1}\Vert \le \Vert u-{\bar{x}}\Vert +\Vert Tx_{n-1}-{\bar{x}}\Vert \le \Vert u-{\bar{x}}\Vert +\Vert x_{n-1}-{\bar{x}}\Vert \le 2C_{{\bar{x}}}, \end{aligned}$$
(4.1)

and

$$\begin{aligned} \Vert x_{n-1}-Tx_{n-1}\Vert \le \Vert x_{n-1}-{\bar{x}}\Vert +\Vert Tx_{n-1}-{\bar{x}}\Vert \le 2\Vert x_{n-1}-{\bar{x}}\Vert \le 2C_{{\bar{x}}}. \end{aligned}$$

Again, it follows from (1.6) and \(\delta _n+\alpha _n+\beta _n=1\) that:

$$\begin{aligned} \Vert x_{n+1}-x_n\Vert&=\Vert \alpha _n(x_n-x_{n-1})+\beta _n(Tx_n-Tx_{n-1})+(\delta _n-\delta _{n-1})(u-Tx_{n-1})\nonumber \\&\quad +(\alpha _n-\alpha _{n-1})(x_{n-1}-Tx_{n-1})\Vert \nonumber \\&\le \alpha _n\Vert x_n-x_{n-1}\Vert +\beta _n\Vert Tx_n-Tx_{n-1}\Vert +|\delta _n-\delta _{n-1}| \ \Vert u-Tx_{n-1}\Vert \nonumber \\&\quad +|\alpha _n-\alpha _{n-1}| \ \Vert x_{n-1}-Tx_{n-1}\Vert \nonumber \\&\le (\alpha _n+\beta _n)\Vert x_n-x_{n-1}\Vert +|\delta _n-\delta _{n-1}| \cdot 2C_{{\bar{x}}}+|\alpha _n-\alpha _{n-1}| \cdot 2C_{{\bar{x}}}\nonumber \\&=(1-\delta _n)\Vert x_n-x_{n-1}\Vert +(\delta _{n-1}-\delta _n) \cdot 4C_{{\bar{x}}}. \end{aligned}$$
(4.2)

Moreover

$$\begin{aligned}\Vert x_1-x_0\Vert \le \Vert x_1-{\bar{x}}\Vert +\Vert x_0-{\bar{x}}\Vert \le 2C_{{\bar{x}}}.\end{aligned}$$

Hence, the convergence rate for the sequence \(\Vert x_n-x_{n-1}\Vert \) is an immediate result of applying Lemma 4.1 to (4.2) with \(a_n=\Vert x_n-x_{n-1}\Vert , b_n=\delta _n, M=C_{{\bar{x}}}\).

Similar to (4.1), it is easy to obtain that \(\Vert u-Tx_n\Vert \le 2C_{{\bar{x}}}\). The convergence rate for \(\Vert x_n-Tx_n\Vert \) is derived by the following arguments:

$$\begin{aligned} \Vert x_n-Tx_n\Vert&\le \Vert x_n-x_{n+1}\Vert +\Vert x_{n+1}-Tx_n\Vert \\&=\Vert x_{n+1}-x_n\Vert +\Vert \delta _n(u-Tx_n)+\alpha _n(x_n-Tx_n)\Vert \\&\le \Vert x_{n+1}-x_n\Vert +\delta _n\Vert u-Tx_n\Vert +\alpha _n\Vert x_n-Tx_n\Vert , \end{aligned}$$

and we have

$$\begin{aligned} \Vert x_n-Tx_n\Vert&\le \frac{1}{1-\alpha _n}\big (\Vert x_{n+1}-x_n\Vert +\delta _n\Vert u-Tx_n\Vert \big )\\&\le \frac{1}{1-\alpha _n}\bigg ( \frac{8C_{{\bar{x}}}}{n+4}+\frac{2}{n+4}\cdot 2C_{{\bar{x}}}\bigg )\\&=\frac{12C_{{\bar{x}}}}{n+2}. \end{aligned}$$

This completes the proof. \(\square \)

Remark 4.3

(1) Obviously, the parameters in Theorem 4.2 satisfy our condition (C9), but do not satisfy Kanzow’s condition (C4).

(2) Theorem 4.2 gives the convergence rate of the generalized Halpern iteration (1.6) using the quantity \(\Vert x_n-Tx_n\Vert \) as a measure of its convergence rate. This convergence rate would imply that an \(\epsilon \)-accuracy solution, in the sense that \(\Vert x_n-Tx_n\Vert \le \epsilon \), is obtainable within no more than \(O(1/\epsilon )\) iterations.

Example 4.4

Continue to consider Example 3.5. Consider the iterative scheme (1.6) with

$$\begin{aligned}\delta _n=\alpha _n=\frac{2}{n+4}, \beta _n=\frac{n}{n+4} \quad \text {and} \quad r_n\equiv 0, \forall n\ge 0.\end{aligned}$$

Now, take \(u=y\in K\), and let \(x_0=u\) be the starting point. Since \(r_n\equiv 0\), \(\delta _n+\alpha _n+\beta _n=1\), \(x_0\in K\), \(u\in K\), and T maps into the convex set K, it follows by induction that the entire sequence \(\{x_n\}\) generated by (1.6) belongs to K. Therefore, we have the representation \(x_n=\lambda _n y\) for \(n\ge 0\) for some \(\lambda _n\in [0,1]\). Taking this into account, we can write (1.6) as

$$\begin{aligned}x_{n+1}=\delta _n u+\alpha _n x_n+\beta _nTx_n+r_n=\frac{2}{n+4} y+\frac{2}{n+4}\lambda _n y.\end{aligned}$$

This implies that

$$\begin{aligned}\Vert x_{n+1}\Vert =\frac{2}{n+4}(1+\lambda _n)\Vert y\Vert .\end{aligned}$$

Hence

$$\begin{aligned}\frac{2\Vert y\Vert }{n+3}\le \Vert x_n-Tx_n\Vert =\Vert x_n\Vert \le \frac{4\Vert y\Vert }{n+3}.\end{aligned}$$

Remark 4.5

It should be noted that the convergence rate proved in Theorem 4.2 means that the sequence \(\{\Vert x_n-Tx_n\Vert \}\) converges to 0 with a rate of O(1/n). However, it can be seen from Example 4.4 that this convergence rate cannot be improved.

5 Application to the split feasibility problem

Let C and Q be the nonempty closed convex subsets of the real Hilbert spaces \(H_1\) and \(H_2\), respectively. Let \(A:H_1\rightarrow H_2\) be a bounded linear operator. The split feasibility problem (SFP) was first introduced by Censor and Elfving [6] to solve the phase retrieval problem, and is formulated as

$$\begin{aligned} \text {Find} \quad x^*\in C \quad \text {and} \quad Ax^*\in Q. \end{aligned}$$
(5.1)

Several iterative algorithms were presented to solve the SFP provided the solution exists; see [16, 18, 19, 31, 34, 35] and references therein. One of the most popular and practical algorithms is the CQ algorithm, which was proposed by Byrne [4]. The CQ algorithm is defined as follows:

$$\begin{aligned} x_{n+1}=P_C(x_n-\gamma A^*(I-P_Q)Ax_n), \end{aligned}$$
(5.2)

where \(\gamma \in (0,\frac{2}{\Vert A\Vert ^2})\), \(A^*\) is the corresponding adjoint operator of A, and \(P_C\) and \(P_Q\) stand for the projections to C and Q, respectively.

Let \(F=A^*(I-P_Q)A\) and \(T=P_C(I-\gamma A^*(I-P_Q)A)=P_C(I-\gamma F)\). Next, we prove that T is a nonexpansive mapping if \(0<\gamma \le \frac{2}{\Vert A\Vert ^2}\). Indeed, from Lemma 2.2, we have

$$\begin{aligned} \langle x-y, Fx-Fy\rangle&=\langle x-y, A^*(I-P_Q)Ax-A^*(I-P_Q)Ay\rangle \\&=\langle Ax-Ay, (I-P_Q)Ax-(I-P_Q)Ay\rangle \\&\ge \Vert (I-P_Q)Ax-(I-P_Q)Ay\Vert ^2\\&\ge \frac{1}{\Vert A\Vert ^2}\Vert A^*(I-P_Q)Ax-A^*(I-P_Q)Ay\Vert ^2\\&=\frac{1}{\Vert A\Vert ^2}\Vert Fx-Fy\Vert ^2. \end{aligned}$$

It follows from Lemma 2.2 and \(0<\gamma \le \frac{2}{\Vert A\Vert ^2}\) that

$$\begin{aligned} \Vert Tx-Ty\Vert ^2&=\Vert P_C(x-\gamma Fx)-P_C(y-\gamma Fy)\Vert ^2\\&\le \Vert (x-y)-\gamma (Fx-Fy)\Vert ^2\\&=\Vert x-y\Vert ^2-2\gamma \langle x-y, Fx-Fy\rangle +\gamma ^2\Vert Fx-Fy\Vert ^2\\&\le \Vert x-y\Vert ^2-2\gamma \frac{1}{\Vert A\Vert ^2} \Vert Fx-Fy\Vert ^2+\gamma ^2\Vert Fx-Fy\Vert ^2\\&=\Vert x-y\Vert ^2-\gamma (\frac{2}{\Vert A\Vert ^2}-\gamma ) \Vert Fx-Fy\Vert ^2\\&\le \Vert x-y\Vert ^2. \end{aligned}$$

This shows that T is a nonexpansive mapping.

Lemma 5.1

[32] Let \(\tau >0\) and \(T=P_C(I-\gamma A^*(I-P_Q)A)\). Then, \({\tilde{x}}\in H_1\) solves the SFP (5.1) if and only if \({\tilde{x}}\) solves the fixed point equation \({\tilde{x}}=T{\tilde{x}}\).

In general, the CQ algorithm has only weak convergence. According to Lemma 5.1, solving the SFP is equivalent to solving the fixed point problem of the nonexpansive mapping T. Therefore, we can prove the following strong convergence result by applying Theorem 3.4.

Theorem 5.2

Assume that the solution set of the SFP, denoted by S, is nonempty. Let \(\{\delta _n\}\), \(\{\alpha _n\}\) and \(\{\beta _n\}\) be real sequences in [0, 1], such that \(\delta _n+\alpha _n+\beta _n\le 1\) for all \(n\ge 0\). Let the sequence \(\{x_n\}\) be generated by

$$\begin{aligned} x_{n+1}=\delta _n u+\alpha _n x_n+\beta _n P_C(x_n-\gamma A^*(I-P_Q)Ax_n)+r_n, \quad n\ge 0, \end{aligned}$$
(5.3)

using an arbitrary starting point \(x_0\in H_1\), where \(u\in H_1\) denotes a fixed element, and \(r_n\) is the residual and \(\gamma \in (0, \frac{2}{\Vert A\Vert ^2}]\). If conditions (C1)-(C3) and (C9) hold, then \(\{x_n\}\) strongly converges to z, where \(z:=P_S u\).

In the following, we present two numerical experiments, to illustrate the performance of the iteration (5.3) with conditions (C1–C3) and (C9). All algorithms in our experiments are coded in MATLAB R2016a on a 16 GB RAM, 1.80 GHz, Intel(R) Core(TM) i7-8565U personal computer. Iter. denotes the numbers of iterations, and CPU denotes the computing time.

It is easy to verify that parameters \(\delta _n=\frac{1}{2(n+2)}, \alpha _n=\frac{1}{2(n+2)}+\frac{1}{4}, \beta _n=\frac{n+1}{n+2}-\frac{1}{4}, r_n\equiv 0\) satisfies conditions (C1–C4), and parameters \(\delta _n=\frac{1}{2(n+2)}, \alpha _n=\frac{1}{2(n+2)}, \beta _n=\frac{n+1}{n+2}, r_n\equiv 0\) satisfies conditions (C1–C3) and (C9). For simplicity, the iteration (5.3) with parameters \(\delta _n=\frac{1}{2(n+2)}, \alpha _n=\frac{1}{2(n+2)}+\frac{1}{4}, \beta _n=\frac{n+1}{n+2}-\frac{1}{4}, r_n\equiv 0\) is abbreviated as Iteration A, and the iteration (5.3) with parameters \(\delta _n=\frac{1}{2(n+2)}, \alpha _n=\frac{1}{2(n+2)}, \beta _n=\frac{n+1}{n+2}, r_n\equiv 0\) is abbreviated as Iteration B.

Here are two examples. The first example solves the SFP in finite-dimensional space, while the other solves it in infinite-dimensional space.

Example 5.3

In this example, we apply the iteration (5.3) to solve the LASSO problem. Let us first recall the well-known LASSO problem [26] which is given as follows:

$$\begin{aligned} \begin{array}{ll} &{}\min \limits _{x\in {\mathbb {R}}^N}\frac{1}{2}\Vert Ax-b\Vert ^2,\\ &{}s.t. \quad \Vert x\Vert _1\le t, \end{array} \end{aligned}$$
(5.4)

where \(A\in {\mathbb {R}}^{M\times N}\), \(b\in {\mathbb {R}}^M\) and \(t>0\) is a given constant. Let \(C=\{x\in {\mathbb {R}}^N \mid \Vert x\Vert _1\le t\}\) and \(Q=\{b\}\), then the minimization problem (5.4) can be seen as an SFP (5.1). Thus, we can apply the iteration (5.3) to solve the problem (5.4).

For the experiments, the matrix A is generated from a normal distribution with mean zero and one variance. The true sparse signal \(x^*\in {\mathbb {R}}^N\) is generated from uniformly distribution in the interval \([-2, 2]\) with random K position nonzero, while the rest is kept zero. The sample data \(b=Ax^*\) (no noise is assumed). The goal is then to recover the K-sparse signal x by solving the LASSO problem (5.4).

In the first test, the parameters are set with \(M=50\), \(N=100\), \(t=K\), \(\gamma =\frac{3}{2\Vert A\Vert ^2}\) and \(u=(0, 0,\cdots , 0)^T\). The stopping criteria are that \(\Vert x_{n+1}-x_n\Vert \le 10^{-6}\). The results are listed in Tables 1, 2, and 3 using different initial points \(x_0\) and K.

Table 1 Numerical results obtained by two iterations (K=5)
Table 2 Numerical results obtained by two iterations (K=10)
Table 3 Numerical results obtained by two iterations (K=20)

Example 5.4

[27] Let \(H_1=H_2=L_2[0,1]\) with the inner product given by

$$\begin{aligned}\langle f,g\rangle =\int _0^1 f(t)g(t)dt.\end{aligned}$$

Let \(C=\{x\in L_2[0,1] \mid \Vert x\Vert _{L_2}\le 1\}\) and \(Q=\{x\in L_2[0,1] \mid \langle x, \frac{t}{2}\rangle =0\}\). Find \(x\in C\), such that \(Ax\in Q\), where \((Ax)(t)=\frac{x(t)}{2}\).

In the second test, we take the stepsize \(\gamma =\frac{1}{\Vert A\Vert ^2}\) and \(u=u(t)=1, \forall t\in [0, 1]\). The stopping criteria are that \(\Vert x_{n+1}-x_n\Vert \le 10^{-6}\). The results are listed in Table 4 using different initial points.

Table 4 Numerical results obtained by two iterations

From Tables 1, 2, 3, and 4, we can see that our parameters can make the algorithm perform better compared with Kanzow’s parameters, in terms of cpu time and the numbers of iterations.