1 Introduction

Throughout this paper, all rings are associative with identity. For a nonempty subset X of a ring R, \(l_R(X)=\{a\in R\mid aX=0\}\) and \(r_R(X)=\{a\in R\mid Xa=0\}\) denote the left and the right annihilators of X in R, respectively. Following A. Hmaimou et al [5], a ring R is left strongly Hopfian if for every endomorphism f of R, the chain ker \(f\subseteq \) ker \(f^2\subseteq \cdots \) stabilizes. Equivalently, R is left strongly Hopfian if the chain of left annihilators \(l_R(a)\subseteq l_R(a^2)\subseteq \cdots \) stabilizes for each \(a\in R\). The class of left strongly Hopfian rings is very large. It contains Noetherian rings, Laskerian rings, rings satisfying acc on d-annihilators, those satisfying acc on d-colons, and so on [4]. If R is a commutative ring, then a left strongly Hopfian ring is also called a strongly Hopfian ring. A. Hmaimou et al [5] showed that for a commutative ring R, the ring R is strongly Hopfian if and only if the polynomial ring R[x] is strongly Hopfian if and only if the Laurent polynomial ring \(R[x;x^{-1}]\) is strongly Hopfian. Let R be a commutative ring. In [4], S. Hizem provided an example of a strongly Hopfian ring R such that the power series ring R[[x]] is not necessarily strongly Hopfian, and also gave some necessary and sufficient conditions for R[[x]] to be strongly Hopfian. For more details and properties of left strongly Hopfian rings, see [2, 4, 5, 7, 8].

Let \(\alpha \) be an endomorphism, and \(\delta \) an \(\alpha \)-derivation of R, that is, \(\delta \) is an additive map such that \(\delta (ab)=\delta (a)b+\alpha (a)\delta (b)\), for a, \(b\in R\). According to Annin [1], a ring R is said to be \(\alpha \)-compatible if for each \(a, b \in R, ab=0 \Leftrightarrow a\alpha (b)=0\). Clearly, this may only happen when the endomorphism \(\alpha \) is injective. Moreover, R is said to be \(\delta \)-compatible if for each \(a, b \in R, ab=0 \Rightarrow a\delta (b)=0\). A ring R is \((\alpha , \delta )\)-compatible if it is both \(\alpha \)-compatible and \(\delta \)-compatible. Recall that a ring R is reversible if \(ab=0\Rightarrow ba=0\) for all a, \(b\in R\), and a ring R is semicommutative if \(ab=0\) implies \(aRb=0\) for any a, \(b\in R\). Clearly, any subring of a reversible ring is also reversible, and if R is a reversible ring, then for any \(n\in \mathbb {N}\) and any permutation \(\sigma \in S_n\), \(x_1x_2\cdots x_n=0\) implies \(x_{\sigma (1)}Rx_{\sigma (2)}R\cdots x_{\sigma (n)}R=0\) for any \(x_i\in R\), \(1\le i\le n\). Reversible rings are semicommutative, but the reverse is not true in general [6, Example 1.5]. Let \(f(x)=a_0+a_1x+\cdots +a_nx^n\in R[x;\alpha ,\delta ]\), \(\{a_0, a_1, \ldots , a_n\}\) denote the subset of R that comprised the coefficients of f(x).

In this article, we are interested in the class of right strongly Hopfian rings and the transfer of this property from an associative ring R to the Ore extension \(R[x;\alpha ,\delta ]\) and the monoid ring R[M]. We first provide some examples of right strongly Hopfian rings. We next show that (1) if R is \((\alpha ,\delta )\)-compatible and \(R[x;\alpha ,\delta ]\) is reversible, then the Ore extension \(R[x;\alpha ,\delta ]\) is right strongly Hopfian if and only if R is right strongly Hopfian; (2) if M is a strictly totally ordered monoid and R[M] a reversible ring, then the monoid ring R[M] is right strongly Hopfian if and only if R is right strongly Hopfian.

2 Extensions of Right Strongly Hopfian Rings

Definition 2.1

A ring R is right strongly Hopfian if the chain of right annihilators \(r_R(a)\subseteq r_R(a^2)\subseteq \cdots \) stabilizes for each \(a\in R\).

The next Lemma is known and very useful. We leave the proof for the reader.

Lemma 2.2

Let \(a\in R\). Then the chain \(r_R(a)\subseteq r_R(a^2)\subseteq \cdots \) stabilizes if and only if there exists \(n>m\) such that \(r_R(a^n)=r_R(a^m)\).

Lemma 2.3

Let \(A\subset B\) be an extension of rings. If B is right strongly Hopfian, then so is A.

Proof

Let \(a\in A\). Then \(r_A(a)=r_B(a)\cap A\).

Proposition 2.4

Let \(T_n(R)\) denote the \(n\times n\) upper triangular matrix ring over a ring R. Then the following conditions are equivalent:

  1. (1)

    R is right strongly Hopfian;

  2. (2)

    \(T_n(R)\) is right strongly Hopfian.

Proof

\((1)\Rightarrow (2)\). Suppose R is right strongly Hopfian and let

$$\begin{aligned} A=\left( \begin{array}{cccc} a_{11}&{}\quad a_{12}&{}\quad \cdots &{}\quad a_{1n}\\ 0&{}\quad a_{22}&{}\quad \cdots &{}\quad a_{2n}\\ \cdots &{}\quad \cdots &{}\quad \cdots &{}\quad \cdots \\ 0&{}\quad 0&{}\quad \cdots &{}\quad a_{nn}\\ \end{array}\right) \in T_n(R). \end{aligned}$$

We proceed by induction on n to show that \(T_n(R)\) is right strongly Hopfian. Let \(n=2\). Put \(\alpha =\left( \begin{array}{ll}a&{}b\\ 0&{}c\\ \end{array}\right) \in T_2(R)\). Since R is right strongly Hopfian, there exists \(m\in \mathbb {N}\) such that for any \(n>m\), \(r_R(a^n)=r_R(a^m)\) and \(r_R(c^n)=r_R(c^m)\). Now we show that \(r_{T_2(R)}(\alpha ^{2m+1})=r_{T_2(R)}(\alpha ^{2m})\). If \(\beta =\left( \begin{array}{ll}x&{}y\\ 0&{}z\\ \end{array}\right) \in r_{T_2(R)}(\alpha ^{2m+1}),\) then

$$\begin{aligned}&\alpha ^{2m+1}\beta \\&\quad ={\left( \begin{array}{ll}a^{2m+1}&{}\quad a^{2m}b+a^{2m-1}bc+\cdots +a^mbc^m+\cdots +bc^{2m}\\ 0&{}\quad c^{2m+1}\\ \end{array}\right) \left( \begin{array}{ll}x&{}y\\ 0&{}z\\ \end{array}\right) }\\&\quad \!=\!{\left( \begin{array}{ll}a^{2m+1}x&{}\quad a^{2m+1}y+(a^{2m}b+a^{2m-1}bc+\cdots +a^mbc^m+\cdots +bc^{2m})z\\ 0&{} \quad c^{2m+1}z\\ \end{array}\right) \!=\!0.}\\ \end{aligned}$$

Thus \(x\in r_R(a^{2m+1})=r_R(a^{2m})\) and \(z\in r_R(c^{2m+1})=r_R(c^{2m})=\cdots =r_R(c^m)\). Hence the equation

$$\begin{aligned} a^{2m+1}y+(a^{2m}b+a^{2m-1}bc+\cdots +a^mbc^m+\cdots +bc^{2m})z=0 \end{aligned}$$

becomes

$$\begin{aligned}&a^{2m+1}y+(a^{2m}b+a^{2m-1}bc+\cdots +a^{m+1}bc^{m-1})z\\&\quad =a^{m+1}(a^my+(a^{m-1}b+a^{m-2}bc+\cdots +bc^{m-1})z)=0.\\ \end{aligned}$$

Then

$$\begin{aligned} a^my+(a^{m-1}b+a^{m-2}bc+\cdots +bc^{m-1})\quad z\in r_R(a^{m+1}), \end{aligned}$$

and so

$$\begin{aligned} a^my+(a^{m-1}b+a^{m-2}bc+\cdots +bc^{m-1})z\in r_R(a^m). \end{aligned}$$

Hence

$$\begin{aligned}&a^m(a^my+(a^{m-1}b+a^{m-2}bc+\cdots +bc^{m-1})z)\\&\quad =a^{2m}y+(a^{2m-1}b+a^{2m-2}bc+\cdots +a^mbc^{m-1})z=0. \end{aligned}$$

Then

$$\begin{aligned}&a^{2m}\beta \\&\quad ={\left( \begin{array}{ll}a^{2m}&{} \quad a^{2m-1}b+a^{2m-2}bc+\cdots +abc^{2m-2}+bc^{2m-1}\\ 0&{} \quad c^{2m}\\ \end{array}\right) \left( \begin{array}{ll}x&{}y\\ 0&{}z\\ \end{array}\right) }\\&\quad ={\left( \begin{array}{ll}a^{2m}x&{} \quad a^{2m}y+(a^{2m-1}b+a^{2m-2}bc+\cdots +a^mbc^{m-1}+\cdots +bc^{2m-1})z\\ 0&{} \quad c^{2m}z\\ \end{array}\right) }\\&\quad ={\left( \begin{array}{ll}0&{} \quad a^{2m}y+(a^{2m-1}b+a^{2m-2}bc+\cdots +a^mbc^{m-1})z\\ 0&{} \quad 0\\ \end{array}\right) =0.} \end{aligned}$$

Hence \(r_{T_2(R)}(\alpha ^{2m+1})\subseteq r_{T_2(R)}(\alpha ^{2m})\) and so \(r_{T_2(R)}(\alpha ^{2m+1})= r_{T_2(R)}(\alpha ^{2m}).\) Therefore \(T_2(R)\) is right strongly Hopfian.

Next, we assume that the result is true for \(n-1\), \(n>2\), and let

$$\begin{aligned} A=\left( \begin{array}{cccc} a_{11}&{}\quad a_{12}&{}\quad \cdots &{}\quad a_{1n}\\ 0&{}\quad a_{22}&{}\quad \cdots &{}\quad a_{2n}\\ \cdots &{}\quad \cdots &{}\quad \cdots &{}\quad \cdots \\ 0&{}\quad 0\quad &{}\cdots &{}\quad a_{nn}\\ \end{array}\right) \in T_n(R). \end{aligned}$$

We show that \(r_{T_n(R)}(A)\subseteq r_{T_n(R)}(A^2)\subseteq \cdots \) stabilizes. Put

$$\begin{aligned} A=\left( \begin{array}{cccc} a_{11}&{}\quad a_{12}&{}\quad \cdots &{}\quad a_{1n}\\ 0&{}\quad a_{22}&{}\quad \cdots &{}\quad a_{2n}\\ \cdots &{}\quad \cdots &{}\quad \cdots &{}\quad \cdots \\ 0&{}\quad 0&{}\quad \cdots &{}\quad a_{nn}\\ \end{array}\right) =\left( \begin{array}{ll}A_{n-1}&{} B\\ 0&{}a_{nn}\\ \end{array}\right) . \end{aligned}$$

By the induction hypothesis, we can find \(m\in \mathbb {N}\) such that for any \(s>m\), \(r_{T_{n-1}(R)}(A_{n-1}^s)=r_{T_{n-1}(R)}(A_{n-1}^m)\) and \(r_R(a_{nn}^s)=r_R(a_{nn}^m)\). Then using the same way as above, we can show that \(r_{T_n(R)}(A^{2m+1})=r_{T_n(R)}(A^{2m})\) and so \(T_n(R)\) is right strongly Hopfian by induction.

\((2)\Rightarrow (1)\) This follows easily from Lemma 2.3.

Corollary 2.5

Let \(L_n(R)\) denote the lower triangular matrix ring over R. Then the following conditions are equivalent:

  1. (1)

    R is right strongly Hopfian;

  2. (2)

    \(L_n(R)\) is right strongly Hopfian.

Let

$$\begin{aligned} S_n(R)= & {} \left. \left\{ \left( \begin{array}{llll} a&{}a_{12}&{}\cdots &{} a_{1n}\\ 0&{}a&{}\cdots &{} a_{2n}\\ \cdots &{}\cdots &{}\cdots &{}\cdots \\ 0&{}0&{}\cdots &{}a\\ \end{array} \right) \right| a, a_{ij}\in R\right\} ,\\ G_n(R)= & {} \left. \left\{ \left( \begin{array}{llll} a_1&{}a_{2}&{}\cdots &{} a_{n}\\ 0&{}a_1&{}\cdots &{} a_{n-1}\\ \cdots &{}\cdots &{}\cdots &{}\cdots \\ 0&{}0&{}\cdots &{}a_1\\ \end{array} \right) \right| a_{i}\in R, 1\le i\le n \right\} , \end{aligned}$$

and let \(R\bowtie R\) denote the trivial extension of R by R.

Corollary 2.6

The following conditions are equivalent:

  1. (1)

    R is right strongly Hopfian;

  2. (2)

    \(S_n(R)\) is right strongly Hopfian;

  3. (3)

    \(G_n(R)\) is right strongly Hopfian;

  4. (4)

    \(R[x]/(x^n)\) is right strongly Hopfian;

  5. (5)

    \(R\bowtie R\) is right strongly Hopfian.

Proof

Note that \(R[x]/(x^n)\cong G_n(R)\) and \(R\bowtie R\cong G_2(R).\)

Let R be a ring. Immediately, we deduce that the lower triangular matrix ring over R is right strongly Hopfian if and only if the upper triangular matrix ring is right strongly Hopfian. Let R be a ring, and let \(W(R)=\left. \left\{ \left( \begin{array}{lll}a_{11}&{}0&{}0\\ a_{21}&{}a_{22}&{}a_{23}\\ 0&{}0&{}a_{33}\\ \end{array}\right) \right| a_{ij}\in R\right\} \). Then W(R) is a \(3\times 3\) subring of \(M_3(R)\) under usual matrix addition and multiplication. A natural problem asks if the right strongly Hopfian property of such a ring coincides with that of R. This inspires us to consider the right strongly Hopfian property of W(R).

Proposition 2.7

Let R be a ring. Then W(R) is right strongly Hopfian if and only if R is right strongly Hopfian.

Proof

Suppose R is right strongly Hopfian and let

$$\begin{aligned} \alpha =\left( \begin{array}{lll}a&{}\quad 0&{}\quad 0\\ x&{} \quad b&{} \quad y\\ 0&{}\quad 0&{} \quad c\\ \end{array}\right) \in W(R). \end{aligned}$$

Then there exists \(m\in \mathbb {N}\) such that for any \(n>m\), \(r_R(a^n)=r_R(a^m)\), \(r_R(b^n)=r_R(b^m)\), and \(r_R(c^n)=r_R(c^m)\). Now we show that \(r_{W(R)}(\alpha ^{2m+1})=r_{W(R)}(\alpha ^{2m})\). If

$$\begin{aligned} \beta =\left( \begin{array}{lll}d&{} \quad 0&{} \quad 0\\ s&{}\quad e&{} \quad t\\ 0&{}\quad 0&{} \quad f\\ \end{array}\right) \in r_{W(R)}(\alpha ^{2m+1}), \end{aligned}$$

then

$$\begin{aligned} \alpha ^{2m+1}\beta =\left( \begin{array}{lll}a^{2m+1}d&{}\quad 0&{} \quad 0\\ ud+b^{2m+1}s&{}\quad b^{2m+1}e&{} \quad b^{2m+1}t+vf\\ 0&{}\quad 0&{} \quad c^{2m+1}f\\ \end{array}\right) =0, \end{aligned}$$

where

$$\begin{aligned} u=xa^{2m}+bxa^{2m-1}+\cdots +b^mxa^m+b^{m+1}xa^{m-1}+\cdots +b^{2m-1}xa+b^{2m}x, \end{aligned}$$

and

$$\begin{aligned} v=b^{2m}y+b^{2m-1}yc+b^{2m-2}yc^2+\cdots +byc^{2m-1}+yc^{2m}. \end{aligned}$$

Hence

$$\begin{aligned} d\in r_R(a^{2m+1})= & {} r_R(a^{2m})=\cdots =r_R(a^m),\\ e\in r_R(b^{2m+1})= & {} r_R(b^{2m})=\cdots =r_R(b^m), \end{aligned}$$

and

$$\begin{aligned} f\in r_R(c^{2m+1})=r_R(c^{2m})=\cdots =r_R(c^m). \end{aligned}$$

Then

$$\begin{aligned} 0= & {} ud+b^{2m+1}s\\= & {} (xa^{2m}+bxa^{2m-1}+\cdots +b^mxa^m+b^{m+1}xa^{m-1}+\cdots +b^{2m}x)d+b^{2m+1}s\\= & {} (b^{m+1}xa^{m-1}+b^{m+2}xa^{m-2}+\cdots +b^{2m}x)d+b^{2m+1}s\\= & {} b^{m+1}((xa^{m-1}+bxa^{m-2}+\cdots +b^{m-1}x)d+b^ms), \end{aligned}$$

and

$$\begin{aligned} 0= & {} b^{2m+1}t+vf\\= & {} b^{2m+1}t+(b^{2m}y+b^{2m-1}yc+\cdots +b^{m+1}yc^{m-1}+b^myc^m+\cdots +yc^{2m})f\\= & {} b^{2m+1}t+(b^{2m}y+b^{2m-1}yc+\cdots +b^{m+1}yc^{m-1})f\\= & {} b^{m+1}(b^mt+(b^{m-1}y+b^{m-2}yc+\cdots +yc^{m-1})f). \end{aligned}$$

Hence

$$\begin{aligned} (xa^{m-1}+bxa^{m-2}+\cdots +b^{m-1}x)d+b^ms\in r_R(b^{m+1})=r_R(b^m) \end{aligned}$$

and

$$\begin{aligned} b^mt+(b^{m-1}y+b^{m-2}yc+\cdots +yc^{m-1})f\in r_R((b^{m+1})=r_R(b^m). \end{aligned}$$

So

$$\begin{aligned}&b^m((xa^{m-1}+bxa^{m-2}+\cdots +b^{m-1}x)d+b^ms)\\&\quad =b^mxa^{m-1}+b^{m+1}xa^{m-2}+\cdots +b^{2m-1}x)d+b^{2m}s=0, \end{aligned}$$

and

$$\begin{aligned}&b^m(b^mt+(b^{m-1}y+b^{m-2}yc+\cdots +yc^{m-1})f)\\&\quad =b^{2m}t+(b^{2m-1}y+b^{2m-2}yc+\cdots +b^myc^{m-1})f=0. \end{aligned}$$

Then by routine computations, we can show that \(\alpha ^{2m}\beta =0\) and so \(\beta \in r_{W(R)}(\alpha ^{2m})\). Hence \(r_{W(R)}(\alpha ^{2m+1})=r_{W(R)}(\alpha ^{2m}).\) Therefore W(R) is right strongly Hopfian.

Conversely, if W(R) is right strongly Hopfian, then by Lemma 2.3 R is right strongly Hopfian.

Let \(\alpha \) be an endomorphism and \(\delta \) an \(\alpha \)-derivation of R. We denote by \( R[x;\alpha , \delta ]\) the Ore extension whose elements are the polynomials over R, the addition is defined as usual, and the multiplication is subject to the relation \(xa=\alpha (a)x+\delta (a)\) for any \(a\in R\). From this rule, an inductive argument can be made in order to calculate an expression for \(x^ja\), for all \(j\in \mathbb {N}\) and \(a\in R\). To recall this result, we shall use some convenient notation introduced in [9].

Notation 2.8

Let \(\delta \) be an \(\alpha \)-derivation of R. For integers ij with \(0\le i\le j\), \(f^j_i\in End(R, + )\) will denote the map which is the sum of all possible words in \(\alpha , \delta \) built with i letters \(\alpha \) and \(j-i\) letters \(\delta \). For instance, \(f_0^0=1, f^j_j=\alpha ^j, f^j_0=\delta ^j\), and \(f^j_{j-1}=\alpha ^{j-1}\delta +\alpha ^{j-2}\delta \alpha +\cdots +\delta \alpha ^{j-1}.\) Using recursive formulas for the \(f_i^j\)’s and induction, as done in [9], one can show with a routine computation that

$$\begin{aligned} x^ja=\sum \limits _{i=0}^jf^j_i(a)x^i. \end{aligned}$$

The following Lemma is well known and we omit the proof (see [3, Lemma 2.1]).

Lemma 2.9

Let R be an \((\alpha ,\delta )\)-compatible ring. Then we have the following:

  1. (1)

    If \(ab=0\), then \(a\alpha ^n(b) =\alpha ^n(a)b=0\) for all positive integers n.

  2. (2)

    If \(\alpha ^k(a)b=0\) for some positive integer k, then \(ab=0\).

  3. (3)

    If \(ab=0\), then \(\alpha ^n(a)\delta ^m(b)=0=\delta ^ m(a)\alpha ^n(b)\) for all positive integers mn.

  4. (4)

    If \(ab=0\), then \(af^j_i(b)=0\) and \(f_i^j(a)b=0\) for all i, j.

Lemma 2.10

Let R be an \(\alpha \)-compatible ring. If \(\alpha ^{k_1}(a_1)\alpha ^{k_2}(a_2)\cdots \alpha ^{k_n}(a_n)=0\) for some positive integers, then \(a_1a_2\cdots a_n=0\).

Proof

Using induction, for \(n=1\), the result is true by the injectivity of \(\alpha \). Now suppose \(\alpha ^{k_1}(a_1)\alpha ^{k_2}(a_2)\cdots \alpha ^{k_n}(a_n)=0\). Then \(\alpha ^{k_1}(a_1)\alpha ^{k_2}(a_2)\cdots \alpha ^{k_{n-1}}(a_{n-1})a_n=0\), and so \(\alpha ^{k_1}(a_1)\alpha ^{k_2}(a_2)\cdots \alpha ^{k_{n-1}}(a_{n-1}a_n)=0\). Then \(a_1a_2\cdots a_n=0\).

Lemma 2.11

Let R be an \((\alpha ,\delta )\)-compatible ring, \(f(x)=a_0+a_1x+\cdots +a_nx^n\) and \(g(x)=b_0+b_1x+\cdots +b_mx^m\) be two polynomials in \(R[x;\alpha ,\delta ].\) Then we have the following:

  1. (1)

    If for all \(0\le i\le n\) and \(0\le j\le m\), \(a_ib_j=0,\) then \(f(x)g(x)=0\).

  2. (2)

    If R is semicommutative and \(c\in R\) is such that for all \(0\le j\le m\), \(cb_j=0\), then \(cf(x)g(x)=0\).

Proof

(1) We have

$$\begin{aligned} f(x)g(x)= & {} (a_0+a_1x+\cdots +a_nx^n)(b_0+b_1x+\cdots +b_mx^m)\\= & {} \sum \limits _{l=0}^{m+n}\left( \sum \limits _{s+t=l}\left( \sum _{i=s}^na_if^i_s(b_t)\right) \right) x^l. \end{aligned}$$

By Lemma 2.9, \(a_ib_t=0\) implies \(a_if^i_s(b_t)=0\). Thus it is easy to see that \(f(x)g(x)=0\).

(2) Since R is semicommutative, for all \(0\le i\le n\) and \(0\le j\le m\), \(cb_j=0\) implies \(ca_ib_j=0\). Thus by (1) we complete the proof.

For two polynomials f(x) and g(x) in \(R[x;\alpha ,\delta ]\), in order to calculate an expression for \((f(x)+g(x))^n\), for all \(n\in \mathbb {N}\), we denote by \([Q^n_if(x)g(x)]\) the polynomial which is the sum of all possible terms, with each term being a product of i polynomials f(x) and \(n-i\) polynomials g(x). Using this convenient notation, we have \((f(x)+g(x))^n=f(x)^n+[Q^n_{n-1}f(x)g(x)]+[Q^n_{n-2}f(x)g(x)]+\cdots +[Q^n_{1}f(x)g(x)]+g(x)^n\).

Lemma 2.12

Let R be an \((\alpha ,\delta )\)-compatible semicommutative ring, \(ax^r\), \(f(x)=b_0+b_1x+\cdots +b_mx^m\), \(g(x)=c_0+c_1x+\cdots +c_qx^q\) be three polynomials in \(R[x;\alpha ,\delta ]\). If \(c_j\in r_R(a^n)\) for all \(0\le j\le q\), then for any \(p>n\), \([Q_n^p(ax^r)f(x)]g(x)=0\).

Proof

It is easy to check that the coefficients of \([Q^p_n(ax^r)f(x)\)] can be written as sums of monomials of length p in \(f^t_s(a)\) and \(f^v_u(b_j)\), where \(b_j\in \{ b_0, b_1,\ldots , b_m\}\) and \(t\ge s\ge 0\), \(v\ge u\ge 0\) are nonnegative positive integers. Consider each monomial \(f^{t_1}_{s_1}(v_1)f^{t_2}_{s_2}(v_2)\cdots f^{t_p}_{s_p}(v_p)\) where \(v_1\), \(v_2\), \({\ldots }\), . It would contain n letters a. Suppose \(v_{r_1}=v_{r_2}=\cdots =v_{r_n}=a\) for some \(1\le r_1< r_2<\cdots <r_n\le p\). Then we write the monomial \(f^{t_1}_{s_1}(v_1)f^{t_2}_{s_2}(v_2)\cdots f^{t_p}_{s_p}(v_p)\) as

$$\begin{aligned} f^{t_1}_{s_1}(v_1)\cdots f^{t_{r_1}}_{s_{r_1}}(a) f^{t_{r_1+1}}_{s_{r_1+1}}(v_{r_1+1})\cdots f^{t_{r_n-1}}_{s_{r_n-1}}(v_{r_n-1})f^{t_{r_n}}_{s_{r_n}}(a)f^{t_{r_n+1}}_{s_{r_n+1}}(v_{r_n+1})\cdots f^{t_p}_{s_p}(v_p), \end{aligned}$$

where \(v_s\in \{b_0, b_1, \ldots , b_m\}\) if \(s\not \in \{r_1, r_2,\ldots , r_n\}\). For each \(0\le j\le q\), since R is \((\alpha ,\delta )\)-compatible and semicommutative, \(a^nc_j=aa\cdots ac_j=0\) implies

$$\begin{aligned} f^{t_{r_1}}_{s_{r_1}}(a)f^{t_{r_2}}_{s_{r_2}}(a)\cdots f^{t_{r_n}}_{s_{r_n}}(a)c_j=0, \end{aligned}$$

and so

$$\begin{aligned}&f^{t_1}_{s_1}(v_1)\cdots f^{t_{r_1}}_{s_{r_1}}(a) f^{t_{r_1+1}}_{s_{r_1+1}}(v_{r_1+1})\cdots \\&f^{t_{r_n-1}}_{s_{r_n-1}}(v_{r_n-1})f^{t_{r_n}}_{s_{r_n}}(a)f^{t_{r_n+1}}_{s_{r_n+1}}(v_{r_n+1})\cdots f^{t_p}_{s_p}(v_p)c_j=0. \end{aligned}$$

Thus by Lemma 2.11 we complete the proof.

The same idea can be used to prove the following.

Corollary 2.13

Let R be an \((\alpha ,\delta )\)-compatible semicommutative ring, \(ax^r\), \(f(x)=b_0+b_1x+\cdots +b_mx^m\), \(g(x)=c_0+c_1x+\cdots +c_qx^q\) be three polynomials in \(R[x;\alpha ,\delta ]\). If \(c_j\in r_R(a^n)\) for all \(0\le j\le q\), Then we have the following:

  1. (1)

    For any \(p>n+l\), \([Q_{n+l}^p(ax^r)f(x)]g(x)=0\).

  2. (2)

    \(R[x;\alpha ,\delta ](a^{i_1}x^{n_1})R[x;\alpha ,\delta ](a^{i_2}x^{n_2})R[x;\alpha ,\delta ]\cdots (a^{i_k}x^{n_k})R[x;\alpha ,\delta ]g(x)=0\) if \(i_1+i_2+\cdots +i_k\ge n\).

Proposition 2.14

Let R be \((\alpha ,\delta )\)-compatible and \(R[x;\alpha ,\delta ]\) be reversible. Then the following conditions are equivalent:

  1. (1)

    R is right strongly Hopfian;

  2. (2)

    \(R[x;\alpha ,\delta ]\) is right strongly Hopfian.

Proof

\((1)\Rightarrow (2)\) Let \(f(x)=a_0+a_1x+\cdots +a_nx^n\in R[x;\alpha ,\delta ]\). Since R is right strongly Hopfian, there exists \(k\in \mathbb {N}\) such that for all \(l>k\) and all \(0\le i\le n\), \(r_R(a_i^l)=r_R(a_i^k)\). Now we show that \(r_{R[x;\alpha ,\delta ]}(f(x)^{(n+1)k+1})=r_{R[x;\alpha ,\delta ]}(f(x)^{(n+1)k}).\) If

$$\begin{aligned} g(x)=b_0+b_1x+\cdots +b_mx^m\in r_{R[x;\alpha ,\delta ]}(f(x)^{(n+1)k+1}), \end{aligned}$$

then

$$\begin{aligned} 0= & {} f(x)^{(n+1)k+1}g(x)=(a_0+a_1x+\cdots +a_nx^n)^{(n+1)k+1}(b_0+b_1x+\cdots +b_mx^m)\\= & {} a_n\alpha ^n(a_n)\alpha ^{2n}(a_n)\cdots \alpha ^{(n+1)kn}(a_n)\alpha ^{[(n+1)k\!+\!1]n}(b_m)x^{[(n+1)k+1]n+m}\!+\! {\mathrm{lower}}\,\, {\mathrm{terms}}. \end{aligned}$$

Hence

$$\begin{aligned} a_n\alpha ^n(a_n)\alpha ^{2n}(a_n)\cdots \alpha ^{(n+1)kn}(a_n)\alpha ^{[(n+1)k+1]n}(b_m)=0. \end{aligned}$$

By Lemma 2.10, we obtain \(a_n^{(n+1)k+1}b_m=0.\) Hence

$$\begin{aligned} b_m\in r_R\left( a_n^{(n+1)k+1}\right) =r_R\left( a_n^k\right) . \end{aligned}$$

From \(f(x)^{(n+1)k+1}g(x)=0\), we have \(a_n^kf(x)^{(n+1)k+1}g(x)=0\). Then by Lemma 2.11 we obtain

$$\begin{aligned} 0= & {} a_n^kf(x)^{(n+1)k+1}g(x)\\= & {} a_n^k(a_0+a_1x+\cdots +a_nx^n)^{(n+1)k+1} (b_0+b_1x+\cdots +b_mx^m)\\= & {} a_n^k(a_0+a_1x+\cdots +a_nx^n)^{(n+1)k+1}(b_0+b_1x+\cdots +b_{m-1}x^{m-1})\\&+a_n^k(a_0+a_1x+\cdots +a_nx^n)^{(n+1)k+1}b_mx^m\\= & {} a_n^k(a_0+a_1x+\cdots +a_nx^n)^{(n+1)k+1}(b_0+b_1x+\cdots +b_{m-1}x^{m-1})\\= & {} a_n^ka_n\alpha ^n(a_n)\cdots \alpha ^{(n+1)kn}(a_n)\alpha ^{[(n+1)k+1]n}(b_{m-1})x^{[(n+1)k+1]n+m-1}+ {\mathrm{lower}}\,\, {\mathrm{terms}}. \end{aligned}$$

Hence

$$\begin{aligned} a_n^{k+1}\alpha ^n(a_n)\alpha ^{2n}(a_n)\cdots \alpha ^{(n+1)kn}(a_n)\alpha ^{[(n+1)k+1]n}(b_{m-1})=0 \end{aligned}$$

and so

$$\begin{aligned} b_{m-1}\in r_R\left( a_n^{(n+2)k+1}\right) =r_R\left( a_n^k\right) . \end{aligned}$$

Using the same method repeatedly, we obtain

\(b_j\in r_R(a_n^{(n+1)k+1})=r_R(a_n^k)\) for all \(0\le j\le m\).

Consider the polynomial f(x) as the sum of two polynomials \(a_nx^n\) and \(h(x)=a_{n-1}x^{n-1}+a_{n-2}x^{n-2}+\cdots +a_0.\) Then by Corollary 2.13 we obtain

$$\begin{aligned} 0= & {} f(x)^{(n+1)k+1}g(x)=\left( a_nx^n+h(x)\right) ^{(n+1)k+1}g(x)\nonumber \\= & {} (a_nx^n)^{(n+1)k+1}g(x)+\left[ Q_{(n+1)k}^{(n+1)k+1}(a_nx^n)h(x)\right] g(x)+\cdots \nonumber \\&+ \left[ Q_{k}^{(n+1)k+1}(a_nx^n)h(x)\right] g(x) +\left[ Q_{k-1}^{(n+1)k+1}(a_nx^n)h(x)\right] g(x)\nonumber \\&+\cdots +\left[ Q_{1}^{(n+1)k+1}(a_nx^n)h(x)\right] g(x) +h(x)^{(n+1)k+1}g(x)\nonumber \\= & {} \left[ Q_{k-1}^{(n+1)k+1}(a_nx^n)h(x)\right] g(x)+\left[ Q_{k-2}^{(n+1)k+1}(a_nx^n)h(x)\right] g(x)+\cdots \nonumber \\&+\left[ Q_{1}^{(n+1)k+1}(a_nx^n)h(x)\right] g(x)+ h(x)^{(n+1)k+1}g(x). \end{aligned}$$
(1)

Multiplying Eq. (1) on the left side by \((a_nx^n)^{k-1}\), then by Lemma 2.12 and Corollary 2.13, we obtain

$$\begin{aligned} (a_nx^n)^{k-1}h(x)^{(n+1)k+1}g(x)=0. \end{aligned}$$

Since \(R[x;\alpha ,\delta ]\) is reversible, this implies

$$\begin{aligned} \left[ Q_{k-1}^{(n+1)k+1}(a_nx^n)h(x)\right] g(x)h(x)^{k-1}=0. \end{aligned}$$
(2)

Multiplying Eq. (1) on the right side by \(h(x)^{k-1}\), we obtain

$$\begin{aligned}&\left[ Q_{k-2}^{(n+1)k+1}(a_nx^n)h(x)\right] g(x)h(x)^{k-1} +\left[ Q_{k-3}^{(n+1)k+1}(a_nx^n)h(x)\right] g(x)h(x)^{k-1}\nonumber \\&\quad +\cdots +h(x)^{(n+1)k+1}g(x)h(x)^{k-1}=0. \end{aligned}$$
(3)

Multiplying Eq. (3) on the left side by \((a_nx^n)^{k-2}\), we obtain

$$\begin{aligned}&(a_nx^n)^{k-2}\left[ Q_{1}^{(n+1)k+1}(a_nx^n)h(x)\right] g(x)h(x)^{k-1}\nonumber \\&\quad +(a_nx^n)^{k-2}h(x)^{(n+1)k+1}g(x) h(x)^{k-1}=0. \end{aligned}$$
(4)

By equation \((a_nx^n)^{k-1}h(x)^{(n+1)k+1}g(x)=0\) and because \(R[x;\alpha ,\delta ]\) is reversible, it is easy to see that \((a_nx^n)^{k-2}\left[ Q_{1}^{(n+1)k+1}(a_nx^n)h(x)\right] g(x)h(x)^{k-1}=0\). Hence Eq. (4) becomes

$$\begin{aligned} (a_nx^n)^{k-2}h(x)^{(n+1)k+1}g(x) h(x)^{k-1}=0. \end{aligned}$$

Since \(R[x;\alpha ,\delta ]\) is reversible, this implies

$$\begin{aligned} \left[ Q_{k-2}^{(n+1)k+1}(a_nx^n)h(x)\right] g(x)h(x)^{k-1}h(x)^{k-2}=0. \end{aligned}$$

Multiplying Eq. (3) on the right side by \(h(x)^{k-2}\), we obtain

$$\begin{aligned}&\left[ Q_{k-3}^{(n+1)k+1}(a_nx^n)h(x)\right] g(x)h(x)^{k-1}h(x)^{k-2}\\&\quad + \cdots +h(x)^{(n+1)k+1}g(x)h(x)^{k-1}h(x)^{k-2}=0. \end{aligned}$$

Continuing this process yields that

$$\begin{aligned} h(x)^{(n+1)k+1}g(x)h(x)^{k-1}h(x)^{k-2}\cdots h(x)=0, \end{aligned}$$

and so

$$\begin{aligned}&h(x)^{(n+1)k+1+\frac{k(k-1)}{2}}g(x)\\&\quad =(a_0+a_1x+\cdots +a_{n-1}x^{n-1})^{(n+1)k+1+\frac{k(k-1)}{2}}(b_0+ b_1x+\cdots b_mx^m)=0, \end{aligned}$$

since \(R[x;\alpha ,\delta ]\) is reversible. Now by the same way as above, we obtain

$$\begin{aligned} b_j\in r_R\left( a_{n-1}^{(n+1)k+1+\frac{k(k-1)}{2}}\right) =r_R\left( a_{n-1}^{k}\right) \end{aligned}$$

for all \(0\le j\le m\). Using induction on n, we obtain

$$\begin{aligned} b_j\in r_R\left( a_i^{(n+1)k+1}\right) =r_R\left( a_i^{k}\right) \end{aligned}$$

for all \(0\le j\le m\) and \(0\le i\le n\). It is now easy to check that \(f(x)^{(n+1)k}g(x)=0\). Hence \(r_{R[x;\alpha ,\delta ]}(f(x)^{(n+1)k+1})=r_{R[x;\alpha ,\delta ]}(f(x)^{(n+1)k})\). Therefore \(R[x;\alpha ,\delta ]\) is right strongly Hopfian.

\((2)\Rightarrow (1)\) It is trivial.

Corollary 2.15

We have the following:

  1. (1)

    If R is \(\alpha \)-compatible and the skew polynomial ring \(R[x;\alpha ]\) is reversible, then the skew polynomial ring \(R[x;\alpha ]\) is right strongly Hopfian if and only if R is right strongly Hopfian.

  2. (2)

    If R is \(\delta \)-compatible and the differential polynomial ring \(R[x;\delta ]\) is reversible, then the differential polynomial ring \(R[x;\delta ]\) is right strongly Hopfian if and only if R is right strongly Hopfian.

Proof

It is an immediate consequence of Proposition 2.14.

Corollary 2.16

([5, Theorem 5.1]).   Let R be a commutative strongly Hopfian ring, then the polynomial ring R[x] is strongly Hopfian.

Let M be a multiplicative monoid. In the following, e will always stand for the identity of M. Then R[M] will denote the monoid ring over R consisting of all elements of the form \(\sum _{i=1}^n r_ig_i\) with \(r_i\in R\), \(g_i\in M\), \(i=1, 2,\ldots ,n\), where the addition is given naturally and the multiplication is given by

$$\begin{aligned} \left( \sum _{i=1}^n r_ig_i\right) \left( \sum _{j=1}^m s_jh_j\right) =\sum _{i=1}^n\sum _{j=1}^m(r_is_j)(g_ih_j). \end{aligned}$$

Recall that the ordered monoid \((M,\le )\) is a strictly ordered monoid if for any g, \(g'\), \(h\in M\), \(g<g'\) implies that \(gh<g'h\) and \(hg<hg'\).

For two elements \(\alpha \) and \(\beta \) in R[M], in order to calculate an expression for \((\alpha +\beta )^n\), for all \(n\in \mathbb {N}\), we denote by \([Q_i^n\alpha \beta ]\) the sum of all possible terms with each term being a product of i elements \(\alpha \) and \(n-i\) elements \(\beta \). Using this convenient notation, we have \((\alpha +\beta )^n=\alpha ^n+[Q_{n-1}^n\alpha \beta ]+[Q_{n-2}^n\alpha \beta ]+\cdots + [Q_1^n\alpha \beta ]+\beta ^n\).

Lemma 2.17

Let \((M,\le )\) be a strictly totally ordered monoid and R a semicommutative ring, \(\alpha =ag\), \(\beta =b_1h_1+b_2h_2+\cdots +b_nh_n\) and \(\gamma =c_1v_1+c_2v_2+\cdots +c_mv_m\) be three elements in R[M]. If there exists a positive integer \(n\in \mathbb {Z}\) such that \(c_j\in r_R(a^n)\) for all \(1\le j\le m\), then for any \(p>n\), \([Q_n^p\alpha \beta ]\gamma =0\).

Proof

The coefficients of \([Q_n^p\alpha \beta ]\) can be written as sums of monomials of length p in a and \(b_j\), where \(b_j\in \{b_1, b_2,\ldots , b_n\}\). Consider one of such monomials, \(d_1d_2\cdots d_p\), where \(d_i\in \{a, b_1, b_2,\ldots , b_n\}\), \(0\le i\le p\). It would contain n letters a. Suppose \(d_{r_1}=d_{r_2}=d_{r_n}=a\) for some \(1\le r_1<r_2<\cdots <r_n\le p\). Then we can write the monomial as \(d_1d_2\cdots d_{r_1-1}ad_{r_1+1}\cdots d_{r_n-1}ad_{r_n+1}\cdots d_p\). Since R is semicommutative and \(c_j\in r_R(a^n)\) for all \(1\le j\le m\), \(a^nc_j=aa\cdots ac_j=0\) implies

$$\begin{aligned} d_1d_2\cdots d_{r_1-1}ad_{r_1+1}\cdots d_{r_n-1}ad_{r_n+1}\cdots d_pc_j=0 \end{aligned}$$

for all \(1\le j\le m\). Hence each monomial appearing in \([Q_n^p\alpha \beta ]\gamma \) is equal to 0. Therefore \([Q_n^p\alpha \beta ]\gamma =0\).

The same idea can be used to prove the following.

Corollary 2.18

Let \((M,\le )\) be a strictly totally ordered monoid and R a semicommutative ring, \(\alpha =ag\), \(\beta =b_1h_1+b_2h_2+\cdots +b_nh_n\) and \(\gamma =c_1v_1+c_2v_2+\cdots +c_mv_m\) be three elements in R[M]. If there exists a positive integer \(n\in \mathbb {Z}\) such that \(c_j\in r_R(a^n)\) for all \(1\le j\le m\), then for any \(p>n +l\), we have the following:

  1. (1)

    \([Q_{n+l}^p\alpha \beta ]\gamma =0\).

  2. (2)

    \(R[M](ag)^{n_1}R[M](ag)^{n_2}R[M]\cdots (ag)^{n_k}R[M]\gamma R[M]=0\) if \(n_1+n_2+\cdots +n_k\ge n\).

Proposition 2.19

Let M be a strictly totally ordered monoid and R[M] a reversible ring. Then the following conditions are equivalent:

  1. (1)

    R is a right strongly Hopfian ring;

  2. (2)

    R[M] is a right strongly Hopfian ring.

Proof

\((1)\Rightarrow (2)\). Let \(\alpha =a_1g_1+a_2g_2+\cdots +a_ng_n\in R[M]\) with \(g_i<g_j\) if \(i<j\). Since R is right strongly Hopfian, there exists \(k\in \mathbb {N}\) such that for all \(l>k\) and all \(1\le i\le n\), \(r_R(a_i^l)=r_R(a_i^k).\) Now we show that \(r_{R[M]}(\alpha ^{nk+1})=r_{R[M]}(\alpha ^{nk})\). If \(\beta =b_1h_1+b_2h_2+\cdots b_mh_m\in r_{R[M]}(\alpha ^{nk+1})\) with \(h_s<h_t\) if \(s<t\), then

$$\begin{aligned} 0=\alpha ^{nk+1}\beta =(a_1g_1+a_2g_2+\cdots +a_ng_n)^{nk+1}(b_1h_1+b_2h_2+\cdots b_mh_m). \end{aligned}$$

Considering the coefficient of the largest element \(g_n^{nk+1}h_m\) in \(\alpha ^{nk+1}\beta \), we obtain \(a_n^{nk+1}b_m=0\). Hence \(b_m\in r_R(a_n^{nk+1})=r_R(a_n^k)\). From \(\alpha ^{nk+1}\beta =0\), we have

$$\begin{aligned} 0= & {} (a_n^ke)\alpha ^{nk+1}\beta =(a_n^ke)(a_1g_1+a_2g_2+\cdots +a_ng_n)^{nk+1}(b_1h_1+b_2h_2+\cdots b_mh_m)\\= & {} (a_n^ke)(a_1g_1+a_2g_2+\cdots +a_ng_n)^{nk+1}(b_1h_1+b_2h_2+\cdots b_{m-1}h_{m-1})\\&+(a_n^ke)(a_1g_1+a_2g_2+\cdots +a_ng_n)^{nk+1}b_mh_m\\= & {} (a_n^ke)(a_1g_1+a_2g_2+\cdots +a_ng_n)^{nk+1}(b_1h_1+b_2h_2+\cdots b_{m-1}h_{m-1})\\= & {} (a_n^ke)\alpha ^{nk+1}(\beta -b_mh_m). \end{aligned}$$

Considering the coefficient of the largest element \(g_n^{nk+1}h_{m-1}\) in \((a_n^ke)\alpha ^{nk+1}(\beta -b_mh_m)\), we obtain

$$\begin{aligned} b_{m-1}\in r_R\left( a_n^{(n+1)k+1}\right) =r_R\left( a_n^k\right) . \end{aligned}$$

Continuing this process yields that \(b_j\in r_R(a_n^{nk+1})=r_R(a_n^k)\) for all \(1\le j\le m\). Consider the element \(\alpha \) as the sum of two elements \(a_ng_n\) and \(\gamma =a_1g_1+a_2g_2+\cdots +a_{n-1}g_{n-1}\). Then by Lemma 2.17 and Corollary 2.18 we obtain

$$\begin{aligned} 0= & {} \alpha ^{nk+1}\beta =(a_ng_n+\gamma )^{nk+1}\beta \nonumber \\= & {} (a_ng_n)^{nk+1}\beta +\left[ Q_{nk}^{nk+1}(a_ng_n)\gamma \right] \beta + \cdots +\left[ Q_{k-1}^{nk+1}(a_ng_n)\gamma \right] \beta +\cdots +\gamma ^{nk+1}\beta \nonumber \\= & {} \left[ Q_{k-1}^{nk+1}(a_ng_n)\gamma \right] \beta +\left[ Q_{k-2}^{nk+1}(a_ng_n)\gamma \right] \beta +\cdots +\gamma ^{nk+1}\beta . \end{aligned}$$
(5)

Multiplying Eq. (5) on the left side by \((a_ng_n)^{k-1}\), and by Corollary 2.18, we obtain \((a_ng_n)^{k-1}\gamma ^{nk+1}\beta =0\) and so \([Q_{k-1}^{nk+1}(a_ng_n)\gamma ]\beta \gamma ^{k-1}=0\) since R[M] is reversible. Multiplying Eq. (5) on the right side by \(\gamma ^{k-1}\), we obtain

$$\begin{aligned} \left[ Q_{k-2}^{nk+1}(a_ng_n)\gamma \right] \beta \gamma ^{k-1}+\left[ Q_{k-3}^{nk+1}(a_ng_n)\gamma \right] \beta \gamma ^{k-1} +\cdots +\gamma ^{nk+1}\beta \gamma ^{k-1}=0.\nonumber \\ \end{aligned}$$
(6)

Multiplying Eq. (6) on the left side by \((a_ng_n)^{k-2}\), we obtain

$$\begin{aligned} (a_ng_n)^{k-2}\left[ Q_1^{nk+1}(a_ng_n)\gamma \right] \beta \gamma ^{k-1}+(a_ng_n)^{k-2}\gamma ^{nk+1}\beta \gamma ^{k-1}=0. \end{aligned}$$

Since R[M] is reversible, \((a_ng_n)^{k-1}\gamma ^{nk+1}\beta =0\) implies

$$\begin{aligned} (a_ng_n)^{k-2}\left[ Q_1^{nk+1}(a_ng_n)\gamma \right] \beta \gamma ^{k-1}=0. \end{aligned}$$

Hence we obtain \((a_ng_n)^{k-2}\gamma ^{nk+1}\beta \gamma ^{k-1}=0\), and so \([Q_{k-2}^{nk+1}(a_ng_n)\gamma ]\beta \gamma ^{k-1}\gamma ^{k-2}=0\) since R[M] is reversible. Then multiplying Eq. (6) on the right side by \(\gamma ^{k-2}\), we obtain

$$\begin{aligned}&\left[ Q_{k-3}^{nk+1}(a_ng_n)\gamma \right] \beta \gamma ^{k-1}\gamma ^{k-2} +\left[ Q_{k-4}^{nk+1}(a_ng_n)\gamma \right] \beta \gamma ^{k-1}\gamma ^{k-2}\\&\quad +\cdots +\gamma ^{nk+1}\beta \gamma ^{k-1}\gamma ^{k-2}=0. \end{aligned}$$

Continuing this process, we obtain \(\gamma ^{nk+1}\beta \gamma ^{\frac{k(k-1)}{2}}=0\) and so \(\gamma ^{nk+1+\frac{k(k-1)}{2}}\beta =0\). Using the same way as above, we can show that

$$\begin{aligned} b_j\in r_R(a_{n-1}^{nk+1+\frac{k(k-1)}{2}})=r_R(a_{n-1}^k) \end{aligned}$$

for all \(1\le j\le m\). Using induction on n, we obtain

$$\begin{aligned} b_j\in r_R\left( a_i^{nk+1}\right) =r_R\left( a_i^k\right) \end{aligned}$$

for all \(1\le j\le m\) and \(1\le i\le n\). Then it is easy to check that \(\alpha ^{nk}\beta =0\). Hence \(\beta \in r_{R[M]}(\alpha ^{nk})\), and so \(r_{R[M]}(\alpha ^{nk+1})=r_{R[M]}(\alpha ^{nk})\). Therefore R[M] is right strongly Hopfian.

\((2)\Rightarrow (1)\) It is trivial.

Corollary 2.20

([5, Corollary 5.4]) Let R be a commutative strongly Hopfian ring, then \(R[x,x^{-1}]\) is strongly Hopfian.