2. Cohomology Theory of Nonassociative Algebras
To avoid misunderstandings we give the necessary definitions.
Definition 1. Let G be a set with a single-valued binary operation (multiplication) , where G satisfies the following conditions:
- (1)
For each a and b in G, there is a unique with and
- (2)
A unique exists, satisfying , which is denoted by and correspondingly,
- (3)
A neutral (i.e., unit) element exists: for each .
The set of all elements commuting and associating with G is
- (4)
,
- (5)
,
- (6)
,
- (7)
,
- (8)
;
is called the center of G.
We call G a metagroup if a set G possesses a single-valued binary operation and satisfies Conditions (1)–(3) and
- (9)
for each a, b, and c in G, where , ;
where shortens a notation , where Ψ denotes a (proper or improper) subgroup of .
Then G will be called a central metagroup if, in addition to , it satisfies the condition
- (10)
for each a and b in G, where .
Particularly, is a left inversion, and is a right inversion.
In view of the nonassociativity of G, in general, a product of several elements of G is usually specified by opening “(” and closing “)” parentheses. We denote the product of elements ,..., in G by , where a vector indicates an order of pairwise multiplications of elements in the row in braces in the following manner. The enumerate positions are as follows: before by 1, between and by 2,..., by n between and , and by after . Then, we put if there are k opening "(" and m closing ")" parentheses in the ordered product at the j-th position of the type , where k and m are nonnegative integers, with and .
Traditionally, denotes the symmetric group of the set . Henceforth, maps and functions on metagroups are assumed to be single-valued unless otherwise specified.
Let be a bijective surjective map satisfying the following condition: for each a and b in G. Then, is called an automorphism of the metagroup G.
Lemma 1. . Let G be a central metagroup. Then, for every ,..., in G, and vectors and indicating an order of pairwise multiplications and , there exists an element such that
- (1)
.
. If G is a metagroup and if v is the neutral element in , then property is satisfied.
Proof. From Conditions (1)–(8) in Definition 1, it follows that itself is a commutative group.
. For , evidently , since for each . For , Formula is a direct consequence of condition in Definition 1. Consider . When u is the identity element of , the statement follows from condition in Definition 1. For any transposition u of two elements of the set , the statement follows from and in Definition 1. Elements of can be obtained by multiplication of pairwise transpositions. Therefore, from the condition , it follows that formula is valid.
Now, let and suppose that this lemma is proved for any products consisting of less than n elements. In view of Properties and in Definition 1, it is sufficient to verify Formula of this lemma for since . In this particular case, . Formula follows from the induction hypothesis, since and hence and putting , where with , .
In the general case, , where j is such that either with and with or with and with . Also, ,...,, ,..., with suitable vectors and . If , then and using the induction hypothesis for and , we get that elements s and t in exist so that , where is a corresponding vector prescribing an order of multiplications.
Again, applying the induction hypothesis to the product of elements , we deduce that there exists , such that
.
Therefore, a case remains when . Let the first multiplication in containing be . We put for each . We also put for each , where either or . Therefore, using previous identities, we rewrite the considered product as with an element of the symmetric group and a vector , indicating an order of pairwise products (see Definition 1). From the induction hypothesis, we deduce that there exists , so that with , because G is the central metagroup. Applying the induction hypothesis for , we infer that exists, such that , where either or , correspondingly. From the induction hypothesis for , it follows that exists, so that , and hence, , where .
. Now, let G be a metagroup and be the neutral element of the symmetric group , where for each . Then, using condition of Definition 1 is unnecessary, because transpositions are already not utilized. For and , we get and , since and for each a and b in G. For , Formula of this lemma follows on from condition in Definition 1. Then, the proof in case by induction is a simplification of that of case . □
Lemma 2. If G is a metagroup, then for each a and , the following identities are fulfilled:
- (1)
;
- (2)
;
- (3)
.
Proof. Conditions – in Definition 1 imply that
- (4)
, ;
- (5)
,
for each a and b in G. Using Condition in Definition 1 and Identities and , we deduce that which leads to .
Let . Then, from Identities and , it follows that which provides .
Now, let . Then, Identities and imply that which demonstrates . □
Definition 2. Let A be an algebra over an associative unital ring , such that A has a natural structure of a -bimodule with a multiplication map , which is right and left distributive, , , and also satisfies the identities , , , , and for any a, b, and c in A, r and s in . Let G be a metagroup and be an associative unital ring.
Henceforth, the ring is assumed to be commutative, unless otherwise specified.
Then, by , a metagroup algebra is denoted over for all formal sums satisfying Conditions below, where n is a positive integer, ,..., are in , and ,..., belong to G:
- (1)
for each s in and a in G,
- (2)
for each s and r in , and ,
- (3)
, , for each a and b in G, .
Note 1. Let M be an additive commutative group such that M is a two-sided G-module, where G is a metagroup. We remind the reader that this means that automorphisms and of M correspond to each . For short, we use and for each .
Note that, usually, M has a natural structure of a two-sided -module, because M is the additive commutative group, where denotes the ring of all integers. Therefore, M is a two-sided G-module if and only if it is a two-sided -module according to the formulas and , where for each .
One can consider the additive group of integers as the trivial two-sided G-module putting for each and , where G is a metagroup.
Example 1. I. Recall the following: Let A be a unital algebra over a commutative associative unital ring F supplied with a scalar involution so that its norm N and trace T maps have values in F and fulfill conditions:
- (1)
with ,
- (2)
with ,
- (3)
for each a and b in A.
If a scalar satisfies the condition , then such element f is called cancelable. For a cancelable scalar f, the Cayley–Dickson doubling procedure provides new algebra over F such that
- (4)
,
- (5)
and
- (6)
for each a and b in A. Then, l is called a doubling generator. From the definitions of T and N, it follows that and . The algebra A is embedded into as , where . This is put by induction , where , , , . Then, are generalized Cayley–Dickson algebras when F is not a field or Cayley–Dickson algebras when F is a field.
It is natural to put , where . If , let be the imaginary part of a Cayley–Dickson number z, and hence , where .
If the doubling procedure starts from , then is a *-extension of F. If has a basis over F with the multiplication table , where and with the involution , , then is the generalized quaternion algebra, and is the generalized octonion (Cayley–Dickson) algebra.
When
and
, each
n by
will denote the real Cayley–Dickson algebra with generators
, such that
,
for each
,
for each
. Note that the Cayley–Dickson algebra
for each
is nonassociative, for example,
, etc. Moreover, for each
, the Cayley–Dickson algebra
is nonalternative (see [
7,
8,
9]). Frequently,
is also denoted by
or
. Then,
is a finite metagroup for each
.
Let be a Cayley–Dickson algebra over a commutative associative unital ring that is characteristically different from two, such that , . Take its basic generators , where . Choose as a multiplicative subgroup contained in the ring , such that for each . Put . Then, is a central metagroup.
II. More generally, let H be a group such that with relations and for each and each h and g in H. Then, is also a metagroup. If the group H is noncommutative, then the latter metagroup can be noncentral (see Condition in Definition 1). Using the notation of Example 1. I, we get that the Cayley–Dickson algebra over the real field with for each n provides an example of a metagroup , where denotes the ring of integers.
III. Generally metagroups need not be central. From given metagroups, new metagroups can be constructed using their direct or semidirect products. Certainly, each group is a metagroup also. Therefore, there are abundant families of noncentral metagroups and also of central metagroups different from groups.
In another way, smashed products of groups and of metagroups can be considered by providing other examples of metagroups (for more detail, see
Section 3).
Definition 3. Let be a ring, which may be nonassociative relative to the multiplication. If the mapping , exists, such that and for each a and b in , m and k in M, then M will be called a generalized left -module or, for short, a left -module or left module over .
If is a unital ring and for each , then M is called a left unital module over , where 1 denotes the unit element in the ring . Symmetrically, a right -module is defined.
If M is a left and right -module, then it is called a two-sided -module or a -bimodule. If M is a left -module and a right -module, then it is called a -bimodule.
A two-sided module M over is called cyclic if an element exists such that and , where and .
Let G be a metagroup. Take a metagroup algebra and a two-sided A-module M, where is an associative unital ring (see Definition 2). Let be a two-sided -module for each , where G is the metagroup. Let M have the decomposition as a two-sided -module. Let M also satisfy the following conditions:
- (1)
and ,
- (2)
and and ,
- (3)
and and
for every in G and and . Then, a two-sided A-module M satisfying conditions – is called smashly G-graded. For short, it is also called "G-graded" instead of "smashly G-graded". In particular, if the module M is G-graded and splits into a direct sum of two-sided -submodules , then we say that that M is directly G-graded. For a nontrivial (nonzero) G-graded module X with the nontrivial metagroup G, it is supposed that exists such that if something else is not outlined.
Similarly, G-graded left and right A-modules are defined. Henceforward, speaking about A-modules (left, right, or two-sided), it is supposed that they are G-graded and, for short, “an A-module” is written instead of “a G-graded A-module”, unless otherwise specified.
If P and N are left A-modules and a homomorphism is such that for each and , then is called a left A-homomorphism. Analogously, right A-homomorphisms are defined for right A-modules. For two-sided A modules, a left and right A-homomorphism is called an A-homomorphism.
For left -modules M and N by , a family of all left -homomorphisms is defined from M into N. A similar notation is used for a family of all -homomorphisms (or right -homomorphisms) of two-sided -modules (or right -modules correspondingly). If an algebra A is specified, a homomorphism may be written for short, instead of an A-homomorphism.
Example 2. Let be a commutative associative unital ring. Also, let G be a metagroup and be a metagroup algebra, where A is considered to be a -algebra. Put , and use induction for each natural number n. Each is supplied with a two-sided A-module structure:
- (1)
, and
and
, and
,
where ;
- (2)
with , (see also Formula in Definition 1 above);
- (3)
with ;
- (4)
with ;
- (5)
where ,
,
, ;
where ,
using the shortened notation;
- (6)
for every in G, where denotes a basic element of over , corresponding to the left ordered tensor product
,
.
Proposition 1. For each metagroup algebra (see Definition 2), an acyclic left A-complex exists.
Proof. Take two-sided A-modules , as in example 2. We construct a boundary -linear operator on . For basic elements, it is given by the following formulas:
- (1)
, where
- (2)
,
- (3)
,...,
- (4)
,
- (5)
,
- (6)
;
- (7)
,...,
- (8)
for each in G. On the other hand, from formulas and in Definition 1, it follows that for each , where
- (9)
,...,
- (10)
for every in G. Therefore, is a left and right A-homomorphism of -modules. In particular, , .
Define a -linear homomorphism , which, for basic elements, has the form
- (11)
for every in G. From Formulas and in Definition 1 and in Lemma 1 the identities
- (12)
- (13)
follow for every element ,..., in metagroup G. Vectors , , and indicate the orders of their multiplication, and . The following identity is evident:
- (14)
for data , and obtained from , , and correspondingly by taking the identity into account for each . Hence, for every in G.
Let be a -linear mapping, such that
- (15)
and for each and . Therefore, from Formulas and , we deduce that is the identity on . Consequently, is a monomorphism.
Therefore, from Formulas , (11), (13), and (14) we infer that , for every ,..., in G (see also Definitions 2 and 3 and the notations above).
Thus, the homotopy conditions
- (16)
for each are fulfilled, where 1 denotes the identity operator on . Therefore, the recurrence relation
- (17)
is accomplished, since .
On the other hand, from Formula , it follows that , as the left A-module, is generated by . Then, proceeding by induction in n with the help of , we deduce that for each , since according to Formulas and .
An opposite algebra exists. The latter, as an -linear space, is the same, but has the multiplication for each . Let denote the enveloping algebra of A. Apparently, coincides with as a left and right A-module. Hence, the mapping provides the augmentation .
Thus, identity means that the left complex . is acyclic. □
Example 3. For the Cayley–Dickson algebra over a field F of characteristics not equal to two, let , as the (multiplicative) metagroup, consist of all elements with , , where are generators of the Cayley–Dickson algebra , . Then, is the module over , where .
Example 4. For a topological space U, it is possible to consider the module of all continuous mappings from U into , , , which is supplied with the box product topology.
Example 5. If is a measure space, where is a σ-additive measure on a σ-algebra of a set U, for and for each k, it is possible to consider the space of all mappings from U into , where is taken relative to its norm induced by the scalar product , , .
Example 6. For an additive group H, one can consider the trivial action of A on H. Therefore, the direct product becomes an A-module for an A-module M. In particular, H may be a ring.
Example 7. If there is another ring and a homomorphism , then each left (or right) -module M can be considered as a left (or right, correspondingly) -module by the rule (or correspondingly) for each and .
Vice versa, if M is a right (or left) -module, then the right (or left, correspondingly) module exists (or , correspondingly), which is called the right (or left correspondingly) covariant -extension of M. Similarly, the contravariant right and left extensions or are defined for right or left -modules M, respectively.
This also can be applied to a metagroup algebra over a commutative associative unital ring as in Example 1. Then, by changing a ring, we get right or and left or algebras over . Then, imposing the relation for each and provides a metagroup algebra over , which also has a two-sided -module structure. It will be denoted by or , respectively. Particularly, this is applicable to cases when or is an embedding.
Notation 1. Let be a metagroup algebra (see Definition 2). Put , and by induction, for each natural number n.
If N is a two-sided A-module, it can also be considered as a left -module by the rule for each , , and , is an enveloping algebra, where denotes the opposite algebra of A, where in corresponds to y in element A.
Theorem 1. If is an acyclic left A-complex for a metagroup algebra , as in Proposition 1, and M is a two-sided A-module satisfying Conditions in Definition 3, then a co-chain complex exists:
- (1)
such that is a co-cycle, if and only if f is a -linear derivation from A into M.
Proof. Notation 1 and Example 2 permit each basic element of over to be written as
- (1)
and
- (2)
, where is a basic element in for every in G, , , .
Each homomorphism is characterized by its values on elements , where ,.., belong to a metagroup G. Consider f as a -linear function from into M. Since M satisfies Conditions in Definition 3, then f has the decomposition
- (2)
,
where for every g and in G.
Therefore, the restrictions follow from Conditions in Definition 3, which take into account the nonassociativity of G:
- (3)
,
- (4)
,
- (5)
for every g and in G, where coefficients are prescribed by Formula in Definition 1. Also,
- (6)
and
- (7)
.
For and , naturally, the identities are fulfilled:
- (8)
, and .
A co-boundary operator exists that takes into account the nonassociativity of the (multiplicative) metagroup G:
- (9)
, where
- (10)
, ;
- (11)
, ;...;
- (12)
; ;
- (13)
, ; with .
From onto , the homomorphism is extended by the -linearity. On the other hand, Condition (1) in Definition 3 implies that
- (14)
For each , exists, so that and , where is the left multiplication operator on b:
- (15)
for every in G. Moreover, (or ) in for and , if and only if , since G is a metagroup.
By virtue of Proposition 1, these formulas imply that for each n, since for every in G. Thus, the complex given by formula is exact.
Particularly, is a co-cycle if and only if
- (16)
for each .
We mention that is isomorphic with M.
The one-dimensional co-chain is determined by the mapping . Taking Formula into account, we infer that it is a co-cycle if and only if
- (17)
for each x and y in G. That is, f is a derivation from the metagroup G into the G-module M. There is the embedding of into A as , since . Thus, f has a -linear extension to a -linear derivation from A into M by the following formula:
- (18)
.
□
Remark 1. Suppose that the conditions of Theorem 1 are fulfilled. A two-dimensional co-chain is a 2-co-cycle, if and only if .
That is,
- (1)
for each and in G.
Usually, denotes the set of all n-co-cycles, and the notation is used for the set of n-co-boundaries in . Since, as the additive group, M is commutative, then there are defined groups of cohomologies as the quotient (additive) groups.
For , the co-boundaries are set as zero, and hence, . In the case , a mapping is a co-boundary if the element exists, for which for each . Such a derivation f is called an inner derivation of A defined by an element . The set of all inner derivations is denoted by .
From the cohomological point of view, the additive group is interpreted as the group of all outer derivations , where ; , where the family of all derivations (-homogeneous derivations) from X into a two-sided module M over is denoted by (or respectively).
A two-co-chain is a two-co-boundary, if a one-co-chain exists such that for each x and y in G, the following identity is fulfilled:
- (2)
,
.
Let be a metagroup algebra over a commutative associative unital ring (also see Definitions 1–3).
Let M, N, and P be left A-modules, and a short exact sequence exists:
- (3)
,
where is an embedding, such that and are left A-homomorphisms. Then, P is called an enlargement of a left A-module M with the help of a left A-module N. If there is another enlargement of M with the help of N,
- (4)
such that an isomorphism exists for which and , then enlargements and are called equivalent, where notates the identity mapping, for each .
It is said that an enlargement clefts, if and only if a left A-homomorphism exists, fulfilling the restriction .
In the particular case when , is also an identifying mapping with the first direct summand, and is a projection on the second direct summand, an enlargement is called trivial.
Theorem 2. Let A be a nonassociative metagroup algebra and let M and N be left A-modules, where , is a metagroup (see Definitions 2 and 3). Then, the family can be supplied with a two-sided A-module structure, such that is the set of classes of modules M with the quotient module N.
Proof. The family evidently has the structure of a left module over a ring (see Definition 3), and it can be supplied with a two-sided A-module structure:
and and :
and .
By virtue of Theorem 1, each element induces a (generalized) derivation by Formula in Theorem 1. Each zero-dimensional co-chain provides an inner derivation due to formula in Theorem 1. Then, one co-cycle f induces an enlargement by Formula in Remark 1 with being the direct sum of left A-modules in which N is a submodule and with the left action of A on N: for each and . Suppose that a class of an one co-cycle f is zero, an element exists so that . Then, elements of the form form its submodule , which is isomorphic with M. Moreover, is the direct sum of A-modules. Thus, an enlargement is trivial.
Vice versa, suppose that an enlargement given by formula in Remark 1 exists. That is, a left A-homomorphism satisfying the restriction exists, where notates the identity mapping on N. It induces , such that for all and for each element a of the algebra A.
Suppose that there is another enlargement which clefts, that is, a left -homomorphism exists, fulfilling the restriction . We put for each ; hence, . Then, is a co-cycle of zero class. □
Theorem 3. Suppose that A is a nonassociative metagroup algebra over a commutative associative unital ring , a left A-module N and a two-sided A-module M are given. Then, is the set of classes of enlargements of A with a kernel M such that and with the quotient algebra A. Moreover, an action of A on M in this enlargement coincides with the structure of a two-sided A-module on M.
Proof. If P is an enlargement with a kernel M such that and a quotient module and with , then and supply M with the two-sided A-module structure. Take a -linear mapping inverse from the left to a natural epimorphism and put for each a and b in A. Then, we infer that and , consequently, .
Taking into account that and for each and , we deduce using Formula in Remark 1 that .
Thus, and hence, with .
It remains to prove that the set S of all elements forms a subalgebra isomorphic with A in P. From the construction of S, it follows that S is a a two-sided -module. We verify that it is closed relative to the multiplication for all a and b in A:
.
If A, M, and f are given, then an enlargement P can be constructed as the direct sum of two-sided -modules and with the multiplication rule for every and in M and and in A. It rests to verify that this multiplication rule is homogeneous over and right and left distributive. At first, we evidently get and for all and and in M and and in A, since and . Moreover, we infer that , and analogously, for all and in M and and in A. □
Definition 4. Let M and P and N be two-sided A-modules, where A is a nonassociative metagroup algebra over a commutative associative unital ring . An A-homomorphism (isomorphism) is called a right (operator) A-homomorphism (isomorphism) if it is such for M and N as right A-modules, that is, and for each x and y in M and (see also Definition 3). If an algebra A is specified, a homomorphism (isomorphism) may be written for short instead of an A-homomorphism (an A-isomorphism, respectively).
An enlargement of M by N is called right inessential if a right isomorphism exists, satisfying the restriction .
Theorem 4. Suppose that M is a two-sided A-module, where A is a nonassociative metagroup algebra over a commutative associative unital ring . Then, for each , a two-sided A-module exists such that is isomorphic with the additive group of equivalence classes of right inessential enlargements of M by .
Proof. Consider two right inessential enlargements and of M by N, where and are embeddings of M into and correspondingly. Take a submodule Q of consisting of all elements satisfying the condition . Then, a quotient module exists, where . Therefore, is isomorphic with M, and homomorphisms and induce a homomorphism of onto N. Hence, the submodule is isomorphic with M. Then, an addition of enlargements is prescribed by the formula . Evidently, sums of equivalent enlargements are equivalent.
For an enlargement of M by N, one takes the direct sum of modules and puts to be its submodule consisting of all elements with , where is an embedding of M into E, . Therefore, a homomorphism induces a homomorphism of onto N, since the mapping is a homomorphism of onto M. Also, the ring is commutative and associative. This induces an enlargement of M by N, denoted by and hence, an operation of scalar multiplication of an enlargement on . From this construction, it follows that equivalent enlargements have equivalent scalar multipliers on .
Let be a -linear span of all elements with ,..., in G such that for each . Next, we put
- (1)
and
- (2)
(also see Notations (2–4) of Proposition 1 and (10)–(13) in Theorem 1) for every in G. That is, is the two-sided A-module, where A has the unit element.
By , we denote the family of all right homomorphisms of into M. For each , let an arbitrary element in the additive group of all n co-chains (that is, n times -linear mappings of A into M) on A with values in M be prescribed by the formula for all in A. Consequently, for each , since for all and g in G. This makes the mapping an -linear isomorphism of onto .
Supply with a two-sided A-module structure
- (3)
and
- (4)
for each and all in G, extending f by -linearity on A from G, where are given by Formulas (10)–(13) in Theorem 1. Thus, the mapping is an operator isomorphism. Consequently, is isomorphic with for each integer (n and p) such that and . On the other hand, is isomorphic with for each ; hence, is isomorphic with .
By virtue of Theorem 2 applied with , we infer that is isomorphic with the additive group of equivalence classes of right inessential enlargements of M by . □
Theorem 5. Let M be a two-sided A-module, where A is a nonassociative metagroup algebra over a commutative associative unital ring . Then, to each -co-cycle , an enlargement of M by a two-sided A-module corresponds such that f becomes a co-boundary in it.
Proof. An -co-cycle induces an enlargement of M by due to Theorem 4. An element h in corresponding to f is characterized by the equality
for all in G. This enlargement as the two-sided A-module is such that . Let for all in A. Therefore, we deduce that .
An n-co-chain , defined by , exists for all in A. Thus, . □
Theorem 6. Let A be a nonassociative metagroup algebra over a commutative associative unital ring . Then, an algebra B over exists such that B contains A and each -homogeneous derivation is the restriction of an inner derivation of B.
Proof. Naturally, an algebra A has the structure of a two-sided A-module. In view of Theorem 1, each derivation of the two-sided algebra A can be considered an element of .
Applying Theorem 5 by induction, one obtains a two-sided A-module Q containing M for which an arbitrary element of is represented as the co-boundary of an element of . At the same time, M and Q satisfy Conditions (1)–(3) in Definition 3. This implies that the natural injection of into maps into zero.
Therefore, a two-sided A-module E exists, which, as a two-sided -module, is a direct sum, , and P is such that for each , an element exists that generally depends on f with the property . The metagroup G corresponds to the algebra A. By enlarging P, if necessary, we can consider that to P, a metagroup G also corresponds in such a manner that properties (1)–(3) in Definition 3 are fulfilled.
Now, we take as the underlying two-sided -module of B and supply it with the multiplication as the semidirect product for each , in A and , in P. An embedding of A into B is for each a in A. This implies that . □
Theorem 7. Suppose that A is a nonassociative metagroup algebra of finite order over a commutative associative unital ring and M is a finitely generated two-sided A-module. Then, M is semisimple if and only if its cohomology group is null for each natural number .
Proof. Certainly, if E is an A-module and N is its A-submodule, then a natural quotient morphism exists. Therefore, an enlargement of a two-sided A-module M by a two-sided A-module N is inessential if and only if there is a submodule T in E complemented to such that T is isomorphic with , where is an embedding of M into E. If M is semisimple, then it is either a simple or a finite product of simple modules, since M is finitely generated. For a finitely generated module E and its submodule N, the quotient module is not isomorphic with E, since the algebra A is of finite order over the commutative associative unital ring .
By virtue of Theorems 4 and 5, if, for an algebra A, its corresponding finitely generated two-sided A-modules are semisimple, then its cohomology groups of dimension are zero.
Vice versa, suppose that for each natural number . Consider a finitely generated two-sided A-module E and its two-sided A-submodule N. At first, we take into account the right A-module structure of E with the same right transformations, but with zero left transformations. Then, the left inessential enlargement of by exists, where is the quotient mapping and is an embedding of into . From Theorem 4, it follows that the enlargement is right inessential. Analogously, considering the left A-module structures and we infer that is also left inessential. □
Note 2. Let A be a nonassociative metagroup algebra over a commutative associative unital ring with a characteristic other than two and three. Its opposite algebra exists. The latter, as an F-linear space, is the same, but with the multiplication for each . To each element or , there is posed a left multiplication operator by the formula or a right multiplication operator for each , respectively. Having the anti-isomorphism operator , , , we get
- (1)
and for each x and h in A. Then, taking into account analogously to formula in Theorem 4, we put
Then, taking into account the multipliers , this gives
- (2)
for all in G. Next, symmetrically, provides the formula for for each and in G. We consider the enveloping algebra . Extending these rules by -linearity on A and from G one supplies the tensor product over with the two-sided A-module structure.
Corollary 1. Let A be a semisimple, nonassociative metagroup algebra of finite order over a commutative associative unital ring with a characteristic other than two and three, and let M be a two-sided A-module described in Note 2. Then, for each natural number .
Proof. Since A is semisimple, then the module M from Note 2 is semisimple. Consequently, the statement of this corollary follows from Theorem 7. □