2 $\alpha$-McCoy Rings and Examples
In this section, we relate the problem on the various McCoy properties of a ring $R$ to an endomorphism $\alpha$ of $R$. We begin with the following definition.
Definition 2.1 An endomorphism $\alpha$ of a ring $R$ is called right (resp., left) McCoy, if for each pair of nonzero polynomials $f(x)=\sum\limits_{i=0}^{m}a_{i}x^{i}$ and $g(x)=\sum\limits_{j=0}^{n}b_{j}x^{j}\in R[x]$ with $\alpha(f(x))g(x)=0$ (resp., $f(x)\alpha(g(x))=0$), there exists a nonzero element $r\in R$ such that $\alpha(f(x))r=0$ (resp.,
$r\alpha(g(x))=0$). A ring $R$ is called right (resp., left)
$\alpha$-McCoy if there exists a right (resp., left) McCoy endomorphism $\alpha$ of $R$. $R$ is an $\alpha$-McCoy ring if it is both right and left $\alpha$-McCoy.
It is clear that every right McCoy ring is right $\alpha$-McCoy. However, we can give the following example to show that there exists a McCoy endomorphism $\alpha$ of a ring $S$ such that $S$ is not a McCoy ring.
Example 2.2 Let $\mathbb{Z}$ be the ring of integers. Consider the ring
$
S = \left\{ {\left( {\begin{array}{*{20}c}
a \;\; b \\
0 \;\; c \\
\end{array} } \right)\left| {a, b, c \in \mathbb{Z}} \right.} \right\}.
$ |
Let $\alpha: S\rightarrow S$ be an endomorphism defined by $\alpha
\left( {\left( {\begin{array}{*{10}c}
a \;\; b \\
0 \;\; c \\
\end{array} } \right)} \right) = \left( {\begin{array}{*{20}c}
a \;\; 0 \\
0 \;\; 0 \\
\end{array} } \right).
$ If $f(x)=\sum\limits_{i=0}^{n} \left(
\begin{array}{cc}
a_{i} \;\; b_{i} \\
0 \;\; c_{i} \\
\end{array}
\right)x^{i}$ and $g(x)=\sum\limits_{j=0}^{m} \left(
\begin{array}{cc}
d_j \;\; e_j \\
0 \;\; f_j \\
\end{array}
\right)x^{j}$ are nonzero polynomials in $S[x]$ such that $\alpha(f(x))g(x)=0$. Then we have
$
\begin{eqnarray*}
\alpha \left( {f\left( x \right)} \right)g\left( x \right)\;\;=\;\;
\sum\limits_{k = 0}^{m + n} {\left( {\sum\limits_{i + j = k}
{\left( {\begin{array}{*{20}c}
{a_i } \;\; 0 \\
0 \;\; 0 \\
\end{array}} \right)\left( {\begin{array}{*{20}c}
{d_j } \;\; {e_j } \\
0 \;\; {f_j } \\
\end{array}} \right)} } \right)} x^k
\\
\;\;=\;\; \sum\limits_{k = 0}^{m + n} {\left( {\sum\limits_{i + j = k} {\left( {\begin{array}{*{20}c}
{a_i \;\; d_j } {a_i \;\; e_j } \\
0 \;\;\;\; 0 \\
\end{array}} \right)} } \right)} x^k = 0.
\end{eqnarray*}
$ |
This implies that
$
\sum\limits_{k = 0}^{n +
m}(\sum\limits_{i+j=k}a_{i}d_{j})x^{k}=0, \sum\limits_{k = 0}^{n +
m}(\sum\limits_{i+j=k}a_{i}e_{j})x^{k}=0.
$ |
Let $f_{1}(x)=\sum\limits_{i=0}^{n}a_{i}x^{i}$,
$g_{1}(x)=\sum\limits_{j=0}^{m}d_{j}x^{j}$ and $g_{2}(x)=\sum\limits_{j=0}^{m}e_{j}x^{j}$. Then we have $f_1(x)g_{1}(x)=f_1(x)g_{2}(x)=0$. Since every reduced ring is an Armendariz ring, it follows that $ a_{i}d_{j}=a_{i}e_{j}=0 $ for each $i, j$. If $a_{i}=0$, then we are done. If $a_{i}\neq 0$, then we have $d_{j}=e_{j}=0$. Now if we let
$
r=\left(\begin{array}{cc} 0 \;\; 0 \\ 0 \;\; f_{j} \\
\end{array}
\right)
$ |
for some $f_{j}\neq 0$, then $r\neq 0$ and $\alpha(f(x))r=0$. This shows that the endomorphism $\alpha$ of $S$ is right McCoy. Similarly, we can prove that the endomorphism $\alpha$ of $S$ is left McCoy. But $S$ is neither left nor right McCoy by [10, Theorem 2.1].
According to [1], an endomorphism $\alpha$ of a ring $R$ is called right (resp., left) reversible if whenever $ab=0$ for $a,
b\in R$, $b\alpha(a)=0$ (resp., $\alpha(b)a=0)$. A ring $R$ is called right (resp., left) $\alpha$-reversible if there exists a right (resp., left) reversible endomorphism $\alpha$ of $R$. $R$ is $\alpha$-reversible if it is both left and right $\alpha$-reversible.
Note 2.3 It is well-known that every reversible ring is a McCoy ring. Based on this fact, one may suspect that every left (resp., right) $\alpha$-reversible ring is McCoy. But this is not true by Example 2.2 and [1, Example 2.2]. In general, we do not know if every $\alpha$-reversible ring is $\alpha$-McCoy. In fact, Example 2.2 shows that a right $\alpha$-reversible ring can be $\alpha$-McCoy.
The next proposition gives more examples of right $\alpha$-McCoy rings.
Proposition 2.4 Let $R$ be a ring and $\alpha$ an endomorphism of $R$. Then $R$ is a right $\alpha$-McCoy ring if and only if $R[x]$ is a right $\alpha$-McCoy ring.
Proof Assume that $R$ is a right $\alpha$-McCoy ring. Let $p(y)=f_{0}+f_{1}y+\cdots+f_{m}y^{m}$,
$q(y)=g_{0}+g_{1}y+\cdots+g_{n}y^{n}$ be in $R[x][y]$ with $\alpha(p(y))q(y)=0$. We also let
$
f_{i}=a_{i_{0}}+a_{i_{1}}x+\cdots+a_{w_{i}}x^{w_{i}},
g_{j}=b_{j_{0}}+b_{j_{1}}x+\cdots+b_{v_{j}}x^{v_{j}}
$ |
for each $0\leq i \leq m$ and $0\leq j \leq n$, where $a_{i_{0}},
a_{i_{1}}, \cdots, a_{w_{i}}, b_{j_{0}}, b_{j_{1}}, \cdots,
b_{v_{j}}\in R$. We claim that $R[x]$ is right $\alpha$-McCoy. Take a positive integer $k$ such that $k>\max\{{\rm deg}(f_{i}),
{\rm deg}(g_{j})\}$ for any $0\leq i \leq m$ and $0\leq j \leq n$, where the degree is as polynomials in $R[x]$ and the degree of zero polynomial is take to be zero. Then
$
p(x^{k})=f_{0}+f_{1}x^{k}+\cdots+f_{m}x^{mk},
q(x^{k})=g_{0}+g_{1}x^{k}+\cdots+g_{n}x^{nk}\in R[x],
$ |
and hence the set of coefficients of the $f_{i}^{, }$s (resp., $g_{j}^{, }$s) equals the set of coefficients of $p(x^{k})$ (resp., $q(x^{k})$). Since $\alpha(p(y))q(y)=0$, we have $\alpha(p(x^{k}))q(x^{k})=0$. It follows that there exists $0\neq r \in R\subseteq R[x]$ such that $\alpha(p(x^{k}))r=0$. This implies that $\alpha(p(y))r=0$, and so $R[x]$ is right $\alpha$-McCoy. Conversely, suppose that $f(y)=\sum\limits_{i=0}^{m}a_{i}y^{i}$,
$g(y)=\sum\limits_{j=0}^{n}b_{j}y^{j} \in R[y] \backslash \{0\}$ such that $\alpha(f(y))g(y)=0$. Since $R[x]$ is right $\alpha$-McCoy, there exists $0\neq r(x)\in R[x]$ such that $\alpha(f(y))r(x)=0$. This shows that $\alpha(a_{i})r(x)=0$ for each $i$. It follows from $0\neq r(x)$ that there exists $0\neq
r_{j}\in R$ such that $\alpha(a_{i})r_{j}=0$ for each $i$. Therefore, $\alpha(f(y))r_{j}=0$ and so $R$ is right $\alpha$-McCoy.
Corollary 2.5 A ring $R$ is a right McCoy ring if and only if $R[x]$ is right McCoy.
Let $R$ be a ring and $\triangle$ a multiplicative monoid in $R$ consisting of central regular elements, and let $\triangle^{-1}R$
=$\{u^{-1}a|u\in \triangle, a\in R\}$, then $\triangle^{-1}R$ is a ring. For an endomorphism $\alpha$ of $R$ with $\alpha(\Delta)\subseteq \Delta$, the induced map $\bar{\alpha}:
\triangle^{-1}R\rightarrow \triangle^{-1}R$ defined by $\bar{\alpha}(u^{-1}a)=\alpha(u)^{-1}\alpha(a)$ is also an endomorphism. We have the following result for the right $\alpha$-McCoy property.
Proposition 2.6 Let $R$ be a ring with an endomorphism $\alpha$. If $R$ is right $\alpha$-McCoy, then $\triangle^{-1}R$ is right $\alpha$-McCoy.
Proof Assume that $R$ is right $\alpha$-McCoy and let
$
f(x)=\sum\limits_{i=0}^{m}u_{i}^{-1}a_{i}x^{i},
g(x)=\sum\limits_{j=0}^{n}v_{j}^{-1}b_{j}x^{j} \in
\bigtriangleup^{-1}R[x]
$ |
with $\alpha(f(x))g(x)=0$. Then we have
$
F(x)=(u_{m}u_{m-1}\cdots u_{0})f(x), G(x)=(v_{n}v_{n-1}\cdots
v_{0})g(x)\in R[x].
$ |
Since $R$ is right $\alpha$-McCoy and $\alpha(F(x))G(x)=0$, this implies that there exists a nonzero $r
\in R$ such that $\alpha(u_{m}u_{m-1}\cdots
u_{0}u_{i}^{-1}a_{i})r=0$ for all $i, j$, and so $\alpha(a_{i})r=0$ since $\triangle$ is a multiplicative monoid in $R$ consisting of central regular elements and $u_{i}, v_{j}\in
\bigtriangleup $ for all $i, j$. It follows that $\alpha(u_{i}^{-1}a_{i})r=\alpha(u_{i})^{-1}\alpha(a_{i})r=0$ for all $i, j$. This shows that $\triangle^{-1}R$ is right $\alpha$-McCoy.
The ring of Laurent polynomials in $x$, with coefficients in a ring $R$, consists of all formal sum $\sum\limits_{i=k}^{n}m_{i}x^{i}$ with obvious addition and multiplication, where $m_{i}\in R$ and $k, n$ are (possibly negative) integers. We denote this ring by $R[x; x^{-1}]$. For an endomorphism $\alpha$ of a ring $R$, the map $\bar{\alpha}: R[x;
x^{-1}]\rightarrow R[x; x^{-1}]$ defined by $\bar{\alpha}(\sum\limits_{i=k}^{n}a_{i}x^{i})=\sum\limits_{i=k}^{n}\alpha(a_{i})x^{i}$ extends $\alpha$ and is also an endomorphism of $R[x; x^{-1}]$.
Corollary 2.7 Let $R$ be a ring. If $R$ is a right $\alpha$-McCoy ring, then $R[x; x^{-1}]$ is right $\alpha$-McCoy.
Proof Let $\triangle=\{{1, x, x^{2}, \cdots}\}$, then clearly $\triangle$ is a multipicatively closed subset of $R[x]$. Since $R[x; x^{-1}]\cong \triangle^{-1}R[x]$, it follows directly from Proposition 2.6 that $R[x; x^{-1}]$ is right $\alpha$-McCoy.
According to [2], an endomorphism $\alpha$ of a ring $R$ is called semicommutative if $ab=0$ implies that $aR\alpha(b)=0$ for all $a, b\in R$. A ring $R$ is called $\alpha$-semicommutative if there exists a semicommutative endomorphism $\alpha$ of $R$. Recall from [3] that a ring $R$ is said to be right linearly McCoy if given nonzero linear polynomials $f(x), g(x)\in R[x]$ with $f(x)g(x)=0$, there exists a nonzero element $r\in R$ with $f(x)r=0$. We can define linearly $\alpha$-McCoy rings similarly. It was proved in [3, Proposition 5.3] that every semicommutative ring is right linearly McCoy. The next example gives an example of right linearly $\alpha$-McCoy rings which is not $\alpha$-semicommutative.
Example 2.8 Let $R=\mathbb{Z}_{2}\oplus
\mathbb{Z}_{2}$, where $\mathbb{Z}_{2}$ is the ring of integers modulo 2. Then $R$ is a right linearly $\alpha$-McCoy ring since $R$ is a commutative reduced ring. Let $\alpha: R\rightarrow
R$ be an endomorphism defined by $\alpha((a, b))=(b, a)$. For $(1,
0), (0, 1)\in R$, we have $(1, 0)(0, 1)=0$ but $(1, 0)(1,
1)\alpha(0, 1)\neq 0$. It follows that $R$ is not $\alpha$-semicommutative.
Let $A(R, \alpha)$ be the subset $\{x^{-i}rx^{i}|r\in R, i\geq 0\}$ of the skew Laurent polynomial ring $R[x, x^{-1}; \alpha]$, where $\alpha: R\rightarrow R$ is an injective ring endomorphism of a ring $R$ (see [5] for more details). Elements of $R[x,
x^{-1}; \alpha]$ are finite sums of elements of the form $x^{-i}rx^{i}$ where $r\in R$ and $i$ is a non-negative integer. Multiplication is subject to $xr=\alpha(r)x$ and $rx^{-1}=x^{-1}\alpha(r)$ for all $r\in R$. Note that for each $j\geq 0$, $x^{-i}rx^{i}=x^{-(i+j)}\alpha^{j}(r)x^{(i+j)}$. It follows that the set $A(R, \alpha)$ of all such elements forms a subring of $R[x, x^{-1}; \alpha]$ with
$
\begin{eqnarray*}
x^{-i}rx^{i}+x^{-j}sx^{j}=x^{-(i+j)}(\alpha^{j}(r)+\alpha^{i}(s))x^{(i+j)}, \\
(x^{-i}rx^{i})(x^{-j}sx^{j})=x^{-(i+j)}(\alpha^{j}(r)\alpha^{i}(s))x^{(i+j)}\end{eqnarray*}
$ |
for $r, s\in R$ and $i, j\geq 0$. Note that $\alpha$ is actually an automorphism of $A(R, \alpha)$.
Proposition 2.9 If $R$ is an $\alpha$-rigid ring, then $A(R, \alpha)$ is right $\alpha$-McCoy.
Proof It follows directly from the fact that $A(R, \alpha)$ is an $\alpha$-rigid ring by [4] and that every $\alpha$-rigid ring is right $\alpha$-McCoy.
Proposition 2.10 Let $R$ be a ring and $\alpha$ an endomorphism of $R$. Then $R$ is a right $\alpha$-McCoy ring if and only if $R[x]/(x^{n})$ is a right $\alpha$-McCoy ring, where $(x^{n}) $ is the ideal generated by $x^{n}$.
Proof Assume that $R$ is right $\alpha$-McCoy and we denote the element $\bar{x}$ in $R[x]/(x^{n})$ by $u$. Then
$
R[x]/(x^{n})=R[u]=R+Ru+\cdots+Ru^{n-1}, $
$ |
where $u$ commutes with elements of $R$ and $u^{n}=0$. Let $f(y)=\sum\limits_{i=0}^{p}f_{i}y^{i}$ and $g(y)=\sum\limits_{j=0}^{q}g_{j}y^{j}$ be nonzero polynomials in $R[u][y]$ with $\alpha(f(y))g(y)=0$, where
$
f_{i}=\sum\limits_{s=0}^{n-1}a_{is}u^{s}, ~~
g_{j}=\sum\limits_{t=0}^{n-1}b_{jt}u^{t}.
$ |
Moreover, if we let $k_{s}(y)=\sum\limits_{i=0}^{p}a_{is}y^{i}$,
$h_{t}(y)=\sum\limits_{j=0}^{q}b_{jt}y^{j}$. Then we have
$
\begin{array}{l}
0\;\; = \;\;\alpha (f(y))g(y) = (\sum\limits_{i = 0}^p \alpha ({f_i}){y^i})(\sum\limits_{j = 0}^q {{g_j}} {y^j})\\
\;\;\;\;\;\; = \;\;(\sum\limits_{i = 0}^p {\sum\limits_{s = 0}^{n - 1} \alpha } ({a_{is}}){u^s}{y^i})(\sum\limits_{j = 0}^q {\sum\limits_{t = 0}^{n - 1} {{b_{jt}}} } {u^t}{y^j})\\
\;\;\;\;\;\; = \;\;\sum\limits_{s = 0}^{n - 1} {(\sum\limits_{i = 0}^p \alpha ({a_{is}}){y^i})} \sum\limits_{t = 0}^{n - 1} {(\sum\limits_{j = 0}^q {{b_{jt}}} {y^j}){u^{s + t}} = (\sum\limits_{s = 0}^{n - 1} \alpha ({k_s}(y)} \sum\limits_{t = 0}^{n - 1} {{h_t}} (y)){u^{s + t}}.
\end{array}
$ |
It follows that $\sum\limits_{s+t=k}\alpha(k_{s}(y))h_{t}(y)=0$, where $k=0, 1, \cdots, n-1$. If $\alpha(k_{0}(y))=0$, take $r=u^{n-1}$. Then we have $0\neq r\in R[u]$, and so
$
\alpha(f(y))r=(\sum\limits_{s=0}^{n-1}(\sum\limits_{i=0}^{p}\alpha(a_{is})y^{i})u^{s})u^{n-1}=(\sum\limits_{i=0}^{p}\alpha(a_{i0})y^{i})u^{n-1}=\alpha(k_{0}(y))u^{n-1}=0.
$ |
If $\alpha(k_{0}(y))\neq 0$, it follows from $g(y)\neq 0$ that there is a minimal $k \in \{0, 1, \cdots n-1 \}$ such that $h_{k}(y)\neq 0$ and $\alpha(k_{0}(y))h_{k}(y)=0$. Since $R$ is right $\alpha$-McCoy, there exists a nonzero element $c\in R$ such that $\alpha(k_{0}(y))c=0$. Let $r'=cu^{n-1}$. Then we have $0\neq
r'\in R[u]$ and
$
\alpha(f(y))r'=(\sum\limits_{s=0}^{n-1}(\sum\limits_{i=0}^{p}\alpha(a_{is})y^{i})u^s)cu^{n-1}
=(\sum\limits_{i=0}^{p}\alpha(a_{i0})y^{i})cu^{n-1}=0.
$ |
Conversely, suppose that
$
f(y)=\sum\limits_{i=0}^{p}a_{i}y^{i},
g(y)=\sum\limits_{j=0}^{q}b_{j}y^{j} \in R[y] \backslash \{0\}
$ |
such that $\alpha(f(y))g(y)=0$. Since $f(y)$ and $g(y)$ are nonzero polynomials of $R[x]/(x^{n})[y]$ and $R[x]/(x^{n})$ is right $\alpha$-McCoy, it follows that there exists $0\neq
r_{1}(x)=\sum\limits_{k=0}^{n-1}c_{k}x^{k} \in R[x]/(x^{n})$ such that $\alpha(f(y))r_{1}(x)=0$. Let $c_{k_{0}}\neq 0$ with $k_{0}$ minimal. Then we obtain $\alpha(f(y))c_{k_{0}}=0$ and so $R$ is right $\alpha$-McCoy.
Corollary 2.11 Let $R$ be a ring and $n$ any positive integer. Then $R$ is right McCoy if and only if $R[x]/(x^{n})$ is right McCoy.
A ring $R$ is called right Ore if given $a, b \in R$ with $b$ regular, there exist $a_{1}, b_{1}\in R$ with $b_{1}$ regular such that $ab_{1}=ba_{1}$. It is well-known that $R$ is a right Ore ring if and only if the classical right quotient ring $Q(R)$ of $R$ exists. Suppose that the classical right quotient ring $Q(R)$ of $R$ exists. Then for an endomorphism $\alpha$ of $R$ and any $ab^{-1} \in Q(R)$ where $a, b \in R$ with $b$ regular, the induced map $\bar{\alpha}:Q(R)\rightarrow Q(R)$ defined by $\bar{\alpha}(ab^{-1})=\alpha(a)\alpha(b)^{-1}$ is also an endomorphism.
Proposition 2.12 Let $R$ be a right Ore ring with $Q(R)$ the classical right quotient ring of $R$. If $\alpha$ is an endomorphism of $R$, then $R$ is right $\alpha$-McCoy if and only if $Q(R)$ is right $\alpha$-McCoy.
Proof Let $F(x)=\sum\limits_{i=0}^{m}\delta_{i}x^{i},
G(x)=\sum\limits_{j=0}^{n}\beta_{j}x^{j}$ be nonzero polynomials in $Q[x]$ with $\alpha(F(x))G(x)=0$. By [6, Proposition 2.1.16], we may assume that $\delta_{i}=a_{i}u^{-1}$ and $\beta_{j}=b_{j}v^{-1}$ with $a_{i}, b_{j}\in R$ for each $i, j$ and regular elements $u, v\in R$. Moreover, for each $j$, there exists $c_{j}\in R$ and a regular element $w\in R$ such that $\alpha(u)^{-1}b_{j}=c_{j}w^{-1}$ also by [6, Proposition 2.1.16]. Let $f(x)=\sum\limits_{i=0}^{m}a_{i}x^{i},
g(x)=\sum\limits_{j=0}^{n}c_{j}x^{j}$. Then we have
$
\begin{eqnarray*}0&\;\;=\;\;&\alpha(F(x))G(x)=(\sum\limits_{i=0}^{m}\alpha(\delta_{i})x^{i})(\sum\limits_{j=0}^{n}\beta_{j}x^{j})\\
&=\;\;&\sum\limits_{k=0}^{m+n}(\sum\limits_{i+j=k}\alpha(a_{i})c_{j}(vw)^{-1})x^{k}
=\alpha(f(x))g(x)(vw)^{-1}. \end{eqnarray*}
$ |
This implies that $\alpha(f(x))g(x)=0$. Then there exists a nonzero $r\in R$ such that $\alpha(f(x))r=0$ since $R$ is a right $\alpha$-McCoy ring. Then $\alpha(a_{i})r=0$ for each $i$, and hence $\alpha(\delta_{i})(\alpha(u)r)=0$. Now $Q$ being right $\alpha$-McCoy follows from the fact that $\alpha(F(x))(\alpha(u)r)=0$ since $\alpha(u)r$ is a nonzero element of $Q$. On the other hand, note that if
$
m(x)=\sum\limits_{i=0}^{m}a_{i}x^{i}, n(x)=\sum\limits_{j=0}^{n}b_{j}x^{j}\in
R[x]
$ |
such that $\alpha(m(x))n(x)=0$. Then there exists a nonzero element $\gamma$ in $Q$ such that $\alpha(m(x))\gamma=0$ since $Q$ is right $\alpha$-McCoy. We may assume $\gamma=d\kappa^{-1}$ with $d$ a nonzero element in $R$ and $\kappa$ a regular element. So we obtain $\alpha(m(x))d\kappa^{-1}=0$, and hence $\alpha(m(x))d=0$. This implies that $R$ is right $\alpha$-McCoy. This completes the proof.
Corollary 2.13 Let $R$ be a right Ore ring and $Q(R)$ be the classical right quotient ring of $R$. Then $R$ is right McCoy if and only if $Q(R)$ is right McCoy.
3 Weak $\alpha$-McCoy Rings and its Properties
Comparing with the definition of a weak McCoy ring, we give the following definition of weak $\alpha$-McCoy rings accordingly.
Definition 3.1 An endomorphism $\alpha$ of a ring $R$ is called right (resp., left) weak McCoy, if for each pair of nonzero polynomials $f(x)=\sum\limits_{i=0}^{m}a_{i}x^{i}$ and $g(x)=\sum\limits_{j=0}^{n}b_{j}x^{j}\in R[x]$ with $\alpha(f(x))g(x)=0$ (resp., $f(x)\alpha(g(x))=0$), there exists a nonzero element $r\in R$ such that $\alpha(a_{i})r \in nil(R)$
(resp., $r\alpha(b_{j})\in nil(R)$). A ring $R$ is called right (resp., left) weak $\alpha$-McCoy if there exists a right (resp., left) weak McCoy endomorphism $\alpha$ of $R$. $R$ is a weak $\alpha$-McCoy ring if it is both right and left $\alpha$-McCoy.
It is clear that every right $\alpha$-McCoy ring is right weak $\alpha$-McCoy. It was shown in [10, Theorem 2.1] that $T_{n}(R)$ is not McCoy for $n\geq 2$. The following example shows that there exists a weak $\alpha$-McCoy endomorphism $\alpha$ of a ring $S$ such that $S$ is not an $\alpha$-McCoy ring.
Example 3.2 Let $R$ be a reduced ring. Consider the ring
$
S = \left\{ {\left(
{\begin{array}{*{20}c}
a&b \\
0&c \\
\end{array} } \right)\left| {a, b, c \in R} \right.} \right\}.
\\
$ |
Let $\alpha: S\rightarrow S$ be an endomorphism defined by
$
\alpha
\left( {\left( {\begin{array}{*{10}c}
a&b \\
0&c \\
\end{array} } \right)} \right) = \left( {\begin{array}{*{20}c}
a &-b \\
0&c \\
\end{array} } \right).
$ |
On the other hand, let
$
f(x)=\left(
\begin{array}{cc}
0&-1 \\
0&0 \\
\end{array}
\right)+\left(
\begin{array}{cc}
1&0 \\
0&0 \\
\end{array}
\right)x, ~~g(x)=\left(
\begin{array}{cc}
0&1 \\
0&0 \\
\end{array}
\right)+\left(
\begin{array}{cc}
0&0 \\
0&-1 \\
\end{array}
\right)x
$ |
be elements in $S[x]$. It is straightforward to check that $\alpha(f(x))g(x)=0$, and we can not find a nonzero element $r \in S$ such that $\alpha(f(x)r=0$. This implies that $S$ is not an $\alpha$-McCoy ring. However, $S$ is a right weak $\alpha$-McCoy ring by the following Proposition 3.3.
Proposition 3.3 Let $R$ be a right weak $\alpha$-McCoy ring and $\alpha$ an endomorphism of $R$. Then $T_{n}(R)$ is a right weak $\alpha$-McCoy ring.
Proof Let $f(x)=\sum\limits_{i=0}^{m}A_{i}x^{i},
g(x)=\sum\limits_{j=0}^{n}B_{j}x^{j}$ be nonzero polynomials in $T_{n}(R)[x]$ with $\alpha(f(x))g(x)=0$, where $A_{i}, B_{j} \in
T_{n}(R)$ for all $i, j$. If we denote by $E_{ij}$ the usual matrix unit with 1 in the $(i, j)$-coordinate and zero elsewhere, then for each $\alpha(A_{i})$ there exists a nonzero element $C=rE_{1n}$ such that $\alpha(A_{i})C \in nil(T_{n}(R))$, where $0\neq r\in R$.
Given a ring $R$ and a bimodule $_RM_R$, the trivial extension of $R$ by $M$ is the ring $T(R, M)=R\bigoplus M$ with the usual addition and the following multiplication
$
(r_1, m_1)(r_2, m_2)=(r_1r_2, r_1m_2+m_1r_2).
$ |
This is isomorphic to the ring of all matrix $\left(%
\begin{array}{cc}
r&m \\
0&r \\
\end{array}%
\right), $ where $r\in R$, $m\in M$ and the usual matrix operations are used. For an endomorphism $\alpha$ of a ring $R$ and the trivial extension $T(R, R)$ of $R$, $\overline{\alpha}$:
$T(R, R)\rightarrow T(R, R)$ defined by
$
\overline \alpha \left( {\left( {\begin{array}{*{20}c}
a \;\; b \\
0 \;\; a \\
\end{array}} \right)} \right) = \left( {\begin{array}{*{20}c}
{\alpha \left( a \right)} \;\; {\alpha \left( b \right)} \\
0 \;\;\;\;\;\; {\alpha \left( a \right)} \\
\end{array}} \right)
$ |
is an endomorphism of $T(R, R)$. Since $T(R, 0)$ is isomorphic to $R$, we can identify the restriction of $\overline{\alpha}$ by $T(R, 0)$ to $\alpha$.
Corollary 3.4 If $R$ is right weak $\alpha$-McCoy, then the trivial extension $T(R, R)$ is a right weak $\alpha$-McCoy ring.
Based on Proposition 3.3, one may suspect that if $R$ is a weak $\alpha$-McCoy ring, then the $n$-by-$n$ full matrix ring $M_{n}(R)$ is weak $\alpha$-McCoy with $n\geq 2$. But the following example erases the possibility.
Example 3.5 Let $R$ be a reduced ring. Then $R$ is a weak $\alpha$-McCoy ring. Put $S=M_{n}(R)$ and let $\alpha$ be an endomorphism of $S$ defined by
$
\alpha
\left( {\left( {\begin{array}{*{10}c}
a \;\; b \\
c \;\; d \\
\end{array} } \right)} \right) = \left( {\begin{array}{*{20}c}
a \;\; -b \\
-c \;\; d \\
\end{array} } \right).
$ |
We also let
$
f(x)=\left(
\begin{array}{cc}
0 \;\; -1 \\
0 \;\;\;\; 0 \\
\end{array}
\right)+\left(
\begin{array}{cc}
1 \;\; 0 \\
0 \;\; 0 \\
\end{array}
\right)x,
~~g(x)=\left(
\begin{array}{cc}
1 \;\; 1 \\
0 \;\; 0 \\
\end{array}
\right)+\left(
\begin{array}{cc}
0 \;\;\;\; 0 \\
-1\;\; -1 \\
\end{array}
\right)x
$ |
be polynomials in $S[x]$. Then we have $\alpha(f(x))g(x)=0$, and it is easy to check that $S$ is not a weak $\alpha$-McCoy ring.
Now we consider the case of direct limits of direct systems of right weak $\alpha$-McCoy rings.
Proposition 3.6 The direct limit of a direct system of right weak $\alpha$-McCoy rings is also right weak $\alpha$-McCoy.
Proof Let $D=\{R_{i}, \phi_{ij}\}$ be a direct system of right weak $\alpha$-McCoy rings $R_{i}$ for $i\in I$ and ring homomorphisms $\phi_{ij}:R_{i}\rightarrow R_{j}$ for each $i\leq
j$ satisfying $\phi_{ij}(1)=1$, where $I$ is a direct partially ordered set. Let $R=\underrightarrow {\lim }R_i$ be the direct limit of $D$ with $\iota_{i}:R_{i}\rightarrow R$ and $\iota_{j}\phi_{ij}=\iota_{i}$. We shall prove that $R$ is a right weak $\alpha$-McCoy ring. Let $\alpha$ be an endomorphism of $R$ and take $x, y\in R$. It follows that $x=\iota_{i}(x_{i}),
y=\iota_{j}(y_{j})$ for some $i, j\in I$ and there is $k\in I$ such that $i\leq k, j\leq k$. Now define $x+y=\iota_{k}(\phi_{ik}(x_{i})+\phi_{jk}(y_{j}))$ and $xy=\iota_{k}(\phi_{ik}(x_{i})\phi_{jk}(y_{j}))$, where $\phi_{ik}(x_{i})$ and $\phi_{jk}(y_{j})$ are in $R_{k}$. It is easy to see that $R$ forms a ring with $0=\iota_{i}(0)$ and $1=\iota_{i}(1)$. Let $\alpha(f(x))g(x)=0$ with $f(x)=\sum\limits_{s=0}^{m}a_{s}x^{s}$ and $g(x)=\sum\limits_{t=0}^{n}b_{t}x^{t}$ in $R[x]$. Then there are $i_{s}, j_{t}, k\in I$ such that $\alpha(a_s)=\iota _{i_s } \left(
{a_{i_s } } \right), b_t = \iota _{j_t } \left( {b_{j_t } }
\right), i_{s}\leq k, j_{t}\leq k$. So we have
$
\alpha(a_{s})b_{t}=\iota_{k}(\phi_{i_sk}(a_{i_s)}\phi_{j_tk}(b_{j_t})),
$ |
and hence
$
\begin{eqnarray*}\alpha(f(x))g(x)&\;\;=\;\;&(\sum\limits_{s=0}^{m}\iota_{k}(\phi_{i_sk}(a_{i_s}))x^{s})
(\sum\limits_{t=0}^{n}\iota_{k}(\phi_{j_tk}(b_{j_t}))x^{t})\\
&=\;\;&
\sum\limits_{d=0}^{m+n}(\sum\limits_{s+t=d}\iota_{k}(\phi_{i_sk}(a_{i_s)}\phi_{j_tk}(b_{j_t})))x^{d}=0\end{eqnarray*}
$ |
in $R_{k}[x]$ since $\alpha(f(x))g(x)=0$. On the other hand, since $R_{k}$ is right weak $\alpha$-McCoy, there exists $s_{k}\in
R_{k}\backslash \{0\}$ such that $\iota_{k}(\phi_{i_sk}(a_{i_s}))s_{k}\in nil(R_{k})$ for all $0\leq i\leq m$. Let $s=\iota_{k}(s_{k})$. Then we have $\alpha(a_{s})s\in nil(R)$ and $R$ is right weak $\alpha$-McCoy.
Corollary 3.7 The direct limit of a direct system of right weak McCoy rings is right weak McCoy.
Proposition 3.8 Let $R$ be a ring and $I$ an ideal of $R$ such that $R/I$ is right weak $\alpha$-McCoy. If $I\subseteq
nil(R)$, then $R$ is a right weak $\alpha$-McCoy ring.
Proof Let $f(x)=\sum\limits_{i=0}^{m}a_{i}x^{i}$ and $g(x)=\sum\limits_{j=0}^{n}b_{j}x^{j}$ be polynomials in $R[x]$ with $\alpha(f(x))g(x)=0$. Then we have $\sum\limits_{i=0}^{m}\alpha(\bar{a}_{i})x^{i})(\sum\limits_{j=0}^{n}\bar{b}_{j}x^{j})=0$ in $R/I$. Since $R/I$ is right weak $\alpha$-McCoy, there exists $n_{i} \in \mathbb{N}$ and $s\not\in I$ such that $(\alpha(\bar{a}_{i})\bar{s})^{n_{i}}=0$. It follows that $(\alpha(a_{i})s)^{n_{i}}\in nil(R)$ since $I\subseteq nil(R)$. This completes the proof.