Introduction

Let \( M \) be some set of finite groups. A group \( G \) is said to be saturated with groups from \( M \), if each finite subgroup of \( G \) lies in a subgroup isomorphic to some element of \( M \).

The main goal of the paper is to prove the following result:

Theorem

Suppose that \( m \) is an integer, \( m\geq 2 \), and \( M \) is a set whose elements are finite simple orthogonal groups of dimension \( n=2m+1 \) over fields of odd characteristic. If \( G \) is a periodic group saturated with groups from \( M \), then \( G \) is isomorphic to a simple orthogonal group \( O_{n}(F) \) for some locally finite field \( F \).


A particular case of this theorem was proved in [1].

1. Preliminary Facts

We will be using the notation and results of [2, 3]. Recall some of them.

Suppose that \( F \) is a field of odd characteristic, \( n=2m+1 \) is an odd integer, \( n\geq 5 \), and \( V \) is a vector space of dimension \( n \) over \( F \), while \( e=\{e_{1},e_{2},\dots,e_{n}\} \) is a basis for \( V \) over \( F \). Let \( f \) be a symmetric bilinear form on \( V \) such that \( f(e_{i},e_{i})=1 \) and \( f(e_{i},e_{j})=0 \) for all \( i,j\in\{1,2,\dots,n\} \), \( i\neq j \). The basis \( e \) is called the standard basis for \( f \). Suppose that \( f_{1} \) is a symmetric bilinear form on \( V \) such that \( f_{1}(e_{i},e_{i})=\mu \), where \( \mu \) is an element of \( F \) which is not a square, while \( f_{1}(e_{i},e_{j})=0 \) for all \( i \) and \( j \) satisfying \( i\neq j \). Then \( f \) and \( f_{1} \) are not isometric, and every nondegenerate form on \( V \) is isometric to \( f \) or \( f_{1} \). The group of linear transformations of \( V \) that preserve \( f \), preserves \( f_{1} \) as well and is denoted by \( GO(V) \). The subgroup \( SO(V)=GO(V)\cap SL(V) \) has index 2 in \( GO(V) \) and is called the special orthogonal group \( V \). The group \( \Omega_{n}(F)=\Omega(V)=[SO(V),SO(V)] \) is simple and \( GO(V)/\Omega(V) \) is an elementary abelian group of order 4. The group \( \Omega_{n}(F)=O_{n}(F) \) is isomorphic to a simple group \( B_{m}(F) \) of Lie type \( B \).

Suppose that \( t \) is an involution (i.e. an element of order 2) from \( L=\Omega(V) \). Then \( V \) is an orthogonal direct sum of the subspaces \( V^{+}(t)=\{v\in V\mid vt=v\} \) and \( V^{-}(t)=\{v\in V\mid vt=-v\} \). Denote the dimensions of \( V^{+}(t) \) and \( V^{-}(t) \) by \( d(t) \) and \( r(t) \). It is clear that \( d(t) \) and \( r(t) \) are the defect and the rank of the transformation \( t-1 \) respectively. Two involutions \( t,t_{1}\in\Omega(V) \) conjugate if and only if \( d(t)=d(t_{1}) \). Therefore, \( \Omega_{n}(F) \) contains exactly \( m=(n-1)/2 \) classes of conjugated involutions.

Lemma 1

Let \( A \) be a maximal elementary abelian subgroup of \( L=\Omega(V) \). Then \( |A|=2^{n-1} \) or \( |A|=2^{n-2} \), \( C_{L}(A)=A \), and each involution from \( L \) is conjugate to an involution from \( A \). If involutions \( t \) and \( t_{1} \) from \( L \) are contained in \( A \) and conjugate in \( L \); i.e., \( d(t)=d(t_{1}) \); then \( t \) and \( t_{1} \) are conjugate in \( N_{L}(A) \).

Let \( t \) be an involution from \( A \). If \( d(t)=1 \), then \( |t^{N_{L}(A)}| \) is equal to \( n \) if \( |A|=2^{n-1} \); and \( 1 \), if \( |A|=2^{n-2} \). If \( d(t)\neq 1 \), then \( |t^{N_{L}(A)}|>n \).

The normalizer \( N_{L}(A) \) in \( L \) acts transitively by conjugation on each of the sets \( D_{s}=\{t\in A\mid d(t)=s\} \), \( s=1,3,\dots \). In the case when \( |A|=2^{n-1} \), the subgroup \( N_{L}(A) \) has a subgroup \( N_{0}\geq A \) such that \( |N_{L}(A):N_{0}|\leq 2 \), \( N_{0}\simeq A:\operatorname{Alt}(n) \) and \( N_{0} \) acts transitively by conjugation on each of the sets \( D_{s} \).

Proof

The subgroup \( A \) is diagonal in some orthogonal basis \( e=\{e_{1},\dots,e_{n}\} \) for the space \( V \); therefore, \( A=A^{*}\cap\Omega(V) \), where \( A^{*}=\{a^{*}\in GO(V)\mid e_{i}a^{*}=\pm e_{i}\} \). Since \( A^{*} \) contains \( -1\not\in\Omega(V) \), the order of \( A \) is at most \( 2^{n-1} \), and we may assume that either the basis \( e \) is standard or \( (e_{i},e_{i})=1 \) for \( i>1 \) and \( (e_{1},e_{1})=\mu \), where \( \mu \) is not a square in \( F \). We start with the second case. Thus, the transformation \( c \), satisfying

$$ e_{1}c=-e_{1},\quad e_{2}c=-e_{2},\quad e_{i}c=e_{i}\text{ for }i>2, $$

does not belong to \( \Omega(V) \); i.e., \( \langle c,-1\rangle\cap A=1 \). On the other hand, every transformation \( a\in A^{*} \) such that \( e_{1}a=e_{1} \) and \( d(a) \) is odd, is contained in \( A \). Hence, \( |A|=2^{n-2} \) and

$$ A=\{a=[1,\alpha_{2},\dots,\alpha_{n}]\mid\alpha_{i}=\pm 1,i=2,\dots,n;\ d(a)\text{ is odd}\}, $$

where \( [\alpha_{1},\alpha_{2},\dots,\alpha_{n}] \) denotes the diagonal matrix with the element \( \alpha_{i} \) in the entry \( (i,i) \), \( i=1,2,\dots,n \). It is clear that the number of elements \( a\in A \) such that \( d(a)=d \) is equal to 1, if \( d=1 \); and

$$ C_{n-1}^{d-1}=\frac{(n-1)(n-2)\dots(n-d+1)}{(d-1)!}, $$

if \( d>1 \). If \( 2\leq i<j<k\leq n \) and \( c=c(i,j,k) \) is a transformation such that \( e_{i}c=e_{j} \), \( e_{j}c=e_{k} \), \( e_{k}c=e_{i} \), and \( e_{l}c=e_{l} \), given \( l\not\in\{i,j,k\} \); then \( [1,\alpha_{2},\dots,\alpha_{n}]^{c}=[1,\beta_{2},\dots,\beta_{n}] \), where \( \beta_{i}=\alpha_{k} \), \( \beta_{j}=\alpha_{i} \), \( \beta_{k}=\alpha_{j} \), and \( \beta_{l}=\alpha_{l} \), given \( l\not\in\{i,j,k\} \). Obviously, \( c\in N_{GO(V)}(A) \) and the order of \( c \) is equal to 3; therefore, \( c\in\Omega(V) \). Conjugating \( a=[1,\alpha_{2},\dots,\alpha_{n}] \) by an element \( g \), equal to the product \( c(i_{1},j_{1},k_{1})\dots c(i_{l},j_{l},k_{l}) \) for a suitable \( i_{s} \), \( j_{s} \), \( k_{s} \), there is no difficulty in obtaining the element \( a^{g}=[1,1,\dots,1,-1,\dots,-1] \), where the first \( d(a) \) of the diagonal elements equal 1, and the rest of them equal \( -1 \). This shows that every two involutions \( t,t_{1}\in A \), that are conjugate in \( L \), i.e. involutions with condition \( d(t)=d(t_{1}) \), are conjugate in \( N_{L}(A) \). Moreover, there exists only one involution \( t \) in \( A \) such that \( d(t)=1 \), and the number of involutions \( t \) with condition \( d(t)=d>1 \) is equal to

$$ C_{n-1}^{d-1}=\frac{(n-1)(n-2)\dots(n-d+1)}{(d-1)!}=\frac{(n-1)(n-2)}{2}\prod\limits_{i=3}^{d-1}\frac{n-i}{i}. $$

We will show that this number is more than \( n \) by induction on \( d\geq 3 \) (recall that \( d \) is odd and \( n\geq 5 \)). For \( d=3 \),

$$ C_{n-1}^{d-1}=\frac{(n-1)(n-2)}{2}=\frac{n^{2}-3n+2}{2}\geq n+1>n. $$

Moreover, \( C_{n-1}^{d-1}=C_{n-1}^{n-d} \); therefore, we may assume that \( d-1\leq n-d \), i.e. \( d\leq\frac{n+1}{2} \). Now, given \( i-1\leq d-1 \), we have \( \frac{n-i}{i}>1 \); i.e., \( \prod\nolimits_{i=3}^{n-1}\frac{n-i}{i}>1 \), which implies that \( C_{n-1}^{d-1}>n \).

Direct checking shows that \( C_{GO(V)}(A)=A^{*} \), and hence \( C_{L}(A)=A \).

Consider the standard basis \( e \). Then

$$ A=\{a=[\alpha_{1},\alpha_{2},\dots,\alpha_{n}]\mid\alpha_{1},\alpha_{2},\dots,\alpha_{n}\in\{1,-1\};\ d(a)\text{ is odd}\}. $$

In particular, \( C_{L}(A)=A \). Clearly, \( N^{*}=N_{GO(V)}(A^{*})=A^{*}:\operatorname{Sym}(n) \) and \( N=N^{*}\cap L=A:H \), where \( H\simeq\operatorname{Sym}(n) \) or \( \operatorname{Alt}(n) \), acts transitively on the set of involutions \( t\in A \) with the common parameter \( d(t) \). Thus, if \( t\in A \) and \( d(t)=d \), then \( |t^{N}|=C_{n}^{d} \). As before, it is an easy check that \( |t^{N}|>n \) if \( d(t)>1 \), and \( |t^{N}|=n \) if \( d(t)=1 \).

Lemma 2

Suppose that \( |F|=q \), while \( A \) is a maximal elementary abelian subgroup \( L=\Omega(V) \) with order \( 2^{n-1} \), and \( a \) and \( b \) are two distinct involutions from \( A \) such that \( d(a)=d(b)=1 \) and \( K=\langle a,b\rangle \). Then \( C_{L}(a)=L_{1}:\langle b\rangle \), where \( L_{1}\simeq\Omega^{\varepsilon}_{n-1}(V) \), \( \varepsilon 1\equiv q(\operatorname{mod}4) \), and \( C_{L}(K)=L_{0}\times K \), where \( L_{0}\simeq\Omega_{n-2}(q) \). Moreover, \( C_{L}(a) \) is maximal in \( L \), and \( C_{L}(K) \) is maximal in \( C_{L}(a) \).

Proof

By Lemma 1, \( A \) contains \( n \) involutions \( t \) for which \( d(t)=1 \) and all of them are conjugate in \( N_{L}(A) \). Moreover, \( N_{L}(A) \) acts double transitively on the set of such involutions; hence, all subgroups \( K \) of the statement of Lemma 2 are conjugate in \( N_{L}(A) \).

Since \( C_{GO(V)}(a)=GO(V^{+})\times GO(V^{-}) \), where

$$ V^{+}=\{v\in V\mid va=v\},\quad V^{-}=\{v\in V\mid va=-a\}; $$

therefore, \( C_{L}(a)=C_{GO(V)}(a)\cap\Omega(V) \) includes \( \Omega(V^{-}) \) as a subgroup of index 2. The proof of Lemma 11.53 in [2] implies that \( \Omega(V^{-})=\Omega^{\varepsilon}(V^{-}) \), where \( \varepsilon \) is defined by a congruence \( q\equiv\varepsilon 1(\operatorname{mod}4) \) and \( a \) is the only involution in \( \Omega(V^{-}) \) such that \( d(a)=1 \); hence, \( b\not\in\Omega(V^{-}) \) and \( C_{L}(a)=\Omega(V^{-})\langle b\rangle \). Further,

$$ C_{L}(K)=C_{C_{L}(a)}(b)=\langle b\rangle\times C_{\Omega(V^{-})}(b)=L_{0}\times K, $$

where \( L_{0}\simeq\Omega_{n-2}(q) \). The maximality of \( C_{L}(K) \) in \( C_{C_{L}(a)}(b) \) and of \( C_{L}(a) \) in \( L \) follows from the maximality of geometric subgroups of \( \Omega_{n}(q) \) and \( \Omega_{n-1}^{\varepsilon}(q) \) (see Tables 3.5.D, E, and F in [4] and Tables 8.31, 33, 39, 50, 52, 58, 66, 68, 74, 82, 84, and 85 in [3].

2. Proof of the Theorem

Let \( n \) be an odd integer, \( n\geq 5 \), while \( G \) is a periodic group such that its every subgroup lies in a subgroup isomorphic to \( \Omega_{n}(q) \) for some odd \( q \) which is a power of a prime. Our goal is to show that \( G \) is isomorphic to \( \Omega_{n}(F) \) for a suitable locally finite field \( F \).

Because \( \Omega_{5}(q)\simeq S_{4}(q) \), according to [5] it is true for \( n=5 \). By induction we may assume that \( n\geq 7 \), and the claim is true when \( n \) is replaced with \( n-1 \).

Let \( M(G) \) be the set of all subgroups of \( G \) isomorphic to elements of \( M \), while \( L=\{\Omega_{n}(q)\mid q\text{ is odd}\} \). If \( M(G) \) has a subgroup isomorphic to \( \Omega_{n}(q) \), where \( q\equiv 1(\operatorname{mod}4) \), then we fix and denote by \( L=L(q) \) one of such subgroups. If there are no such subgroups, then we fix and denote as \( L=L(q) \) some (arbitrary) element of \( M(G) \). In both cases we will identify \( L \) with \( \Omega(V) \), where \( V \) is an orthogonal space of dimension \( n \) over a field of order \( q \).

We will also fix an elementary abelian 2-subgroup \( A \) of order \( 2^{n-1} \) from \( L \), elements \( a \) and \( b \) from \( A \), and a subgroup \( K \) as they are defined in Lemma 2. If \( L_{1} \) is a subgroup of \( G \) isomorphic to \( \Omega(V_{1}) \) for some space \( V_{1} \) of dimension \( n \) over a field of order \( q_{1} \), and \( t \) is an involution from \( L_{1} \); then denote by \( d_{L_{1}}(t) \) the dimension of the space of fixed points of \( t \) in \( V_{1} \).

Lemma 3

\( C_{G}(A)=A \), \( N_{G}(A) \) contains a subgroup of index \( 1 \) or \( 2 \) coinciding with \( N_{0}\simeq A:\operatorname{Alt}(n) \) from Lemma \( 1 \).

Proof

Suppose that \( c\in C_{G}(A) \). Then \( C=\langle c,A\rangle \) is a finite subgroup lying in some element \( L_{1}\in M(G) \). Applying Lemma 1 with \( L_{1} \) instead of \( L \), we get that \( c\in C_{L_{1}}(A)=A \). So, \( C_{G}(A)=A \), and therefore \( N_{G}(A) \) is a finite subgroup lying in some \( L_{2}\in M(G) \). Now we use Lemma 1 with \( L_{2} \) in place of \( L \) and derive that \( N_{G}(A) \) includes \( N_{0} \) as a subgroup of index 1 or 2.

Lemma 4

If \( K\leq L_{1}\in M(G) \), then \( d_{L_{1}}(a)=d_{L_{1}}(b)=d(a)=1 \).

Proof

Let \( A_{1} \) be a maximal elementary abelian 2-subgroup of \( L_{1} \) including \( K \).

If \( A_{1}=A \), Lemma 1 with \( L_{1} \) in place of \( L \) implies that the subgroup \( N_{0} \) lies in \( L_{1} \). Because \( |a^{N_{0}}|=|b^{N_{0}}|=n \), we have \( d_{L_{1}}(a)=d_{L_{1}}(b)=1=d_{L}(a) \). In this case the claim of the lemma is true.

Suppose that \( A_{1}\leq A \) and \( A_{1}\neq A \). Then \( |A:A_{1}|=2 \) by Lemma 1. Set \( C=C_{G}(A_{1}) \). It is clear that \( A\leq C \); and, if \( t\in A\setminus A_{1} \), then

$$ C_{C}(t)=C_{G}(t)\cap C=C_{G}\left(\langle t,A_{1}\rangle\right)=C_{G}(A)=A. $$

By Shunkov’s Theorem [6], \( C \) is locally finite. Let \( N_{1}=N_{L_{1}}(A_{1})\leq N_{G}(A_{1}) \). Since \( A_{1} \) is finite, \( N_{G}(A_{1})/C \) is finite, and so \( N_{G}(A_{1}) \) is a locally finite group. It follows that \( N_{G}(A_{1})=CH \), where \( H \) is a finite subgroup including \( A \). Suppose that \( H\leq L_{2}\in M(G) \).

Because \( A\leq L_{2} \), we have \( d_{L_{2}}(a)=d_{L_{2}}(b)=1 \). Lemma 1 implies that \( 1=|a^{N_{G}(A_{1})}|=|a^{H}| \), which yields that \( d_{L_{1}}(a)=d_{L_{1}}(b)=1 \). On the other hand, by Lemma 1\( A_{1} \) has the only involution \( i \) with condition \( d(i)=1 \); therefore, the case under consideration when \( A\neq A_{1}\leq A \) is impossible.

By induction on \( s=|A:(A_{1}\cap A)|+|A_{1}:(A_{1}\cap A)| \), we will show that \( |A_{1}|=|A|=2^{n-1} \) and \( d_{L_{1}}(a)=d_{L_{1}}(b)=1 \).

If \( s\leq 3 \), then either \( A=A_{1}\cap A \), or \( A_{1}=A_{1}\cap A \), and these cases are already done. Hence, we can assume that \( A_{1}\neq A_{1}\cap A\neq A \).

Let \( t\in A\setminus(A_{1}\cap A) \), \( t_{1}\in A_{1}\setminus(A_{1}\cap A) \). Then \( R=\langle t,t_{1},A_{1}\cap A\rangle \) is finite, and so \( R\leq L_{2}\in M(G) \). Suppose that \( \langle t,A_{1}\cap A\rangle\leq A_{2} \), where \( A_{2} \) is a maximal elementary abelian subgroup in \( L_{2} \). Then

$$ |A_{2}|\geq|A_{2}\cap A|>|A_{1}\cap A|. $$

By the inductive hypothesis, \( |A_{2}|=|A| \) and \( d_{L_{2}}(a)=d_{L_{2}}(b)=d(a)=1 \). Further, \( \langle t_{1},A_{1}\cap A\rangle\leq L_{1}\cap L_{2} \) and \( \langle t_{1},A_{1}\cap A\rangle\leq A_{3} \), where \( A_{3} \) is a maximal elementary abelian 2-subgroup of \( L_{2} \). Because \( a,b\in A_{3} \), we have \( |A_{3}|=|A| \). Moreover, \( |A_{3}\cap A_{1}|>|A\cap A_{1}| \). By the inductive hypothesis, \( d_{L_{1}}(a)=d_{L_{1}}(b)=d_{L_{2}}(a)=1 \). The lemma is proved.

Lemma 5

\( C_{G}(K)=K\times R \), where \( R\simeq\Omega_{n-2}(F) \) for some locally finite field \( F \) of odd characteristic.

Proof

Let \( \overline{C}=C_{G}(K)/K \). We want to show that \( \overline{C} \) is saturated with groups from the set \( M_{1}=\{\Omega_{n-2}(q)\mid q\text{ is odd}\} \).


Suppose that \( \overline{X} \) is a finite group from \( \overline{C} \), while \( X \) is the full preimage of \( \overline{X} \) in \( G \). By condition, \( K\leq X\leq L_{1}\in M(G) \); and by Lemma 2\( C_{L_{1}}(K)=K\times R_{1} \), where \( R_{1}\simeq\Omega_{n-2}(q_{1}) \) for odd \( q_{1} \). Thus, \( \overline{C} \) is saturated with groups from the set \( M_{1}=\{\Omega_{n-2}(q)\mid q\text{ is odd}\} \). By the inductive hypothesis, \( \overline{C}\simeq\Omega_{n-2}(F) \) for some locally finite field \( F \) of odd characteristic. In particular, \( C \) is a locally finite group. We will show that \( [C,C]\cap K=1 \). Let \( c\in[C,C] \). Then \( c=[c_{1},c_{2}][c_{3},c_{4}]\dots[c_{p-1}c_{p}] \) for some \( p \) and suitable elements \( c_{1},\dots,c_{p}\in C \). The subgroup \( \langle K,c_{1},\dots,c_{p}\rangle \) is finite and lies in \( K\times Y \), where \( Y\simeq\Omega_{n-2}(q_{2}) \) for some \( q_{2} \). It is clear that \( c\in Y \) and \( c\not\in K \). Since \( C=K[C,C] \); therefore, \( C=K\times[C,C] \), and the lemma is proved.

Lemma 6

\( C_{G}(K) \) lies in a subgroup \( P \) of \( C_{G}(a) \) which is the union of an ascending sequence of subgroups \( P_{i} \), \( i=1,2,\dots \), isomorphic to \( \Omega_{n-1}^{\lambda}(q_{i}).2 \), with \( q_{i}=\lambda 1\ (\operatorname{mod}4) \), \( \lambda\in\{+,-\} \), and \( \lambda \) depends on the choice of \( L \) and is common for all \( i \).

Proof

By Lemma 5, \( C_{G}(K) \) is locally finite and countable. If \( C_{G}(K) \) is finite, then we may assume that \( C_{G}(K)=C_{L}(K) \), and the lemma is true by Lemma 2. Suppose that \( C_{G}(K) \) is infinite and \( C_{G}(K)=\{g_{i}\mid i\in{𝕅}\} \). Put \( P_{0}=C_{L}(a) \). Let \( g_{i_{1}} \) be an element of \( C_{G}(K) \) not belonging to \( P_{1} \), and the number \( i_{1} \) is the smallest of those subject to that condition. The subgroup \( \langle C_{P_{0}}(K),N_{0}\rangle \) coincides with \( P_{0} \) by Lemma 2. Let \( L_{1} \) be an element of \( M(G) \) containing \( C_{P_{0}}(K) \) and let \( g_{i} \) be the first element in order not belonging to \( C_{P_{0}}(K) \). By condition, \( L_{1}\simeq\Omega_{n}(q_{1}) \) for some \( q_{1} \). The subgroup \( C_{L_{1}}(K) \) is maximal in \( C_{L_{1}}(a) \); and, because \( N_{0}\not\leq C(K) \), the subgroup \( \langle C_{L_{1}}(K),N_{0}\rangle \) coincides with \( C_{L_{1}}(a)\simeq\Omega_{n-1}^{\lambda}(q_{1}).2 \). Since \( C_{L_{0}}(K)<C_{L_{1}}(K) \); therefore, \( P_{0}=C_{L_{0}}(a)<C_{L_{1}}(a)=P_{1} \).

Similarly, let \( L_{2} \) be an element of \( M(G) \) including \( C_{P_{1}}(K) \) and let \( g_{i_{2}} \) be the first element in order not belonging to \( C_{P_{1}}(K) \). As before, \( P_{2}=C_{L_{2}}(a) \), and \( P_{2} \) includes \( C_{L_{1}}(a) \). Proceeding this construction of the subgroups \( P_{i} \)’s in a similar way, we will get an ascending sequence of subgroups \( P_{i}\simeq\Omega_{n-1}^{\lambda}(q_{i}) \) whose union \( P \) includes \( C_{G}(a) \). The lemma is proved.

Lemma 7

\( P=C_{G}(a) \).

Proof

Suppose the contrary. Let \( t\in C_{G}(a)\setminus P \). The subgroup \( \langle K,K^{t}\rangle \) is generated by the elements \( a \), \( b \), and \( b^{t} \). Since \( \langle b,b^{t}\rangle \) is a finite group, so is \( \langle K,K^{t}\rangle \). Because \( \langle K,K^{t}\rangle \) lies in the subgroup \( L^{*} \) isomorphic to \( \Omega_{n}(q^{*}) \) for some \( q^{*} \); therefore,

$$ \langle K,K^{t}\rangle\leq D=C_{L^{*}}(a)\simeq\Omega_{n-1}^{\lambda}(q^{*})\cdot 2. $$

Now, \( K \) and \( K^{t} \) are conjugate in \( D \), because \( A \) and \( A^{t} \) are conjugate in \( D \) and \( N_{D}(A) \) acts double transitively on the set of involutions \( a^{*}\in A \) with condition \( \alpha_{D}(a^{*})=1 \). Let \( \Gamma \) be a graph with vertex set \( \Sigma=\{K^{d}\mid d\in D\} \) such that two vertices \( K^{r} \) and \( K^{s} \) are adjacent if and only if \( [K^{r},K^{s}]=1 \). Suppose that \( \Delta \) is a connected component in \( \Gamma \) that includes \( K \). Since \( K\leq A \) and \( C_{N}(K)\neq C_{N}(a) \), we have \( K^{C_{N}(a)}\neq\{K\} \), and so \( |\Delta|\geq 2 \). Thus, if \( \Delta\neq\Sigma \), then \( D \) acts on \( \Sigma \) by conjugation transitively and imprimitively. Hence, the stabilizer of the vertex \( K \) in \( D \) equal to \( N_{D}(K) \) is not maximal in \( D \), which is untrue. Therefore, \( \Delta=\Sigma \) and there is a sequence \( t_{1},t_{2},\dots,t_{r}=t \) of elements from \( C_{L^{*}}(a) \) such that \( 1=[K,K^{t_{1}}]=[K^{t_{i}},K^{t_{i+1}}] \), \( i=1,2,\dots,r-1 \). By induction on \( r \), we will show that \( K^{t_{r}}\leq P \). If \( r=1 \), then \( K^{t}\leq C_{G}(K) \); and by Lemma 6\( K^{t}\leq P \). Suppose that \( r>1 \) and \( K^{t_{r-1}}\leq P \). There exists \( u\in P \) such that \( K^{t_{1}u}=K \) and \( 1=[K,K^{t_{2}u}]=\dots=[K^{t_{r-1}u},K^{t_{r}u}] \), \( i=2,\dots,r \). By the inductive hypothesis, \( K^{t_{r}u}\leq P \) and \( K^{t_{r}}\leq P \).

So, \( K^{t}\leq P \) for every \( t\in C_{G}(a) \). The subgroup \( P \) is locally finite; therefore, \( \langle K,t\rangle \) is finite and lies in \( L^{*}\in M(G) \). Suppose that \( H=C_{L^{*}}(a) \). Then

$$ H=\langle C_{L^{*}}(K),N_{G}(A)\cap C_{L^{*}}(a)\rangle\leq P. $$

Because \( t\in H \), we have \( t\in P \), and the lemma is proved.

Lemma 8

\( C_{G}(a) \) lies in a subgroup \( Z\simeq\Omega_{n}(F) \) of \( G \) for some locally finite field \( F \).

Proof

By Lemma 7, \( C_{G}(a) \) is countable and locally finite. Let \( L_{0}=L \) and

$$ C_{G}(a)=C_{L_{0}}(a)\cup\{g_{i}\in C_{G}(a)\mid i\in{𝕅}\}. $$

The subgroup \( C_{1}=\langle C_{L}(a),g_{i_{1}}\rangle \), with \( g_{i_{1}} \) the first element in order not belonging to \( L_{0} \), is finite and hence lies in \( L_{1}\in M(G) \). Because \( L_{0} \) includes \( N_{G}(A) \) and \( C_{L_{1}}(a) \) is maximal in \( L_{1} \), the subgroup \( \langle C_{1},N_{G}(A)\rangle \) coincides with \( L_{1} \).

Let \( C_{2}=\langle C_{1},g_{i_{2}}\rangle \), where \( g_{i_{2}} \) is the first element in order which is not contained in \( L_{1} \). By condition, \( C_{2}\leq L_{2}\in M(G) \). It is clear that \( C_{2}\leq C_{L_{2}}(a) \) and \( L_{2} \) includes \( N_{L}(A)=N_{L_{2}}(A) \). Since \( C_{L_{2}}(a) \) is maximal in \( L_{2} \) and \( N_{L_{2}}(A)\not\leq C_{L_{2}}(a) \), the subgroup \( L_{2} \) coincides with \( \langle C_{L_{2}}(a),N_{L_{2}}(A)\rangle \) and includes \( L_{1} \).

Reasoning similarly, we construct subgroups \( L_{3},L_{4},\dots\in M(G) \) with condition \( L_{i}\leq L_{i+1} \), \( i=3,4,\dots \). The union \( Z \) of the so-obtained sequence includes \( C_{G}(a) \). By the main result of each of the papers [7,8,9,10,11], \( Z\simeq\Omega_{n}(F) \) for some locally finite field \( F \), and the lemma is proved.

Lemma 9

\( Z=G \).

Proof

By Lemma 8, \( Z \) is countable and locally finite. Suppose that \( g\in G \) and \( a^{g}\neq a \). The group \( \langle a,a^{g}\rangle \) is finite. Therefore, \( \langle a,a^{g}\rangle \) lies in such a subgroup \( R \) of the group \( G \) that is isomorphic to \( \Omega_{n}(r) \) for some \( r \). Let \( \Delta \) be a set of involutions belonging to \( R \) and conjugate to \( a \) in \( G \). Let \( \Gamma \) be a graph with vertex set \( \Delta \) such that two vertices \( a^{g_{1}} \) and \( a^{g_{2}} \) are adjacent if and only if \( [a^{g_{1}},a^{g_{2}}]=1 \). Since \( C_{R}(a) \) is a maximal subgroup of \( R \) and \( |C_{R}(a)\cap\Delta|\geq 2 \); we have by analogy to Lemma 8 that the graph \( \Gamma \) is connected. This implies as in Lemma 8 that \( a^{G}\subseteq Z \), \( \langle a^{G}\rangle=Z \), and \( Z\trianglelefteq G \). Because \( Z \) is locally finite, \( \langle a,g\rangle \) is finite, and we may assume that \( \langle a,g\rangle\leq R \). Since \( \langle a^{R}\rangle=R \); therefore, \( g\in\langle c^{R}\rangle\leq Z \). The lemma is proved, which completes the proof of the theorem.