This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\MSdates

[‘Received date’]‘Accepted date’

Distribution and Generalized Centre in Planar Nearrings

Tim Boykett
Institute for Algebra
   Johannes Kepler University    4040 Linz    Austria    and
Time’s Up Research
   Industriezeile 33b    4020 Linz    Austria
tim@timesup.org
   tim.boykett@jku.at
Abstract

Nearrings are the nonlinear generalization of rings. Planar nearrings play an important role in nearring theory, both from the structural side, being close to generalized nearfields, as well as from an applications perspective, in geometry and combinatorial designs related to difference families. In this paper we investigate the distributive elements of planar nearrings. If a planar nearring has nonzero distributive elements, then it is an extension of its zero multiplier part by an abelian group. In the case that there are distributive elements that are not zero multipliers, then this extension splits, giving an explicit description of the nearring. This generalizes the structure of planar rings. We provide a family of examples where this does not occur, the distributive elements being precisely the zero multipliers. We apply this knowledge to the question of determining the generalized center of planar nearrings as well as finding new proofs of other older results.

1 Introduction

Nearrings are the nonlinear generalization of rings, having only one distributive law. Planar nearrings are a special class of nearrings that generalize nearfields, themselves a generalization of fields, which play an important role in the structural theory of nearrings[14, 20], as well as having important geometric and combinatorial properties[6, 10, 15].

In this paper we look at the distributive elements of a planar nearring, in some sense the ring-like elements of the planar nearring. This extends Aichinger’s work on planar rings [2]. We find that the general structure of planar nearrings with nontrivial distributive elements generalizes the structure of planar rings. We use this information to then investigate the generalized center of a planar nearrings, building on Farag, Cannon, Kabza and Aichinger’s work [3, 7]. This was the original motivation for this work and can be found in §6. As the results there indicate, it was necessary to obtain a good understanding of the distributive elements of a planar nearring, which we undertake in §4 and find some small applications of in §5. The structure implied by nontrivial distributive elements forces several special forms of planar nearrings, which we introduce in §3 and use repeatedly throughout.

In the next section we introduce the necessary background about nearrings and planar nearrings.

2 Background

A (right) nearring (N,+,)(N,+,*) is an algebra such that

  • (N,+)(N,+) is a group with identity 0

  • (N,)(N,*) is a semigroup

  • for all a,b,cNa,b,c\in N, (a+b)c=ac+bc(a+b)*c=a*c+b*c (right distributivity)

While it is readily seen that 0a=00*a=0 for all aNa\in N, it is not necessary that a0=0a*0=0 for all aa. If this is the case, the nearring is called zero symmetric.

A subset of NN is a subnearring if it is closed as an additive group and a multiplicative semigroup. A subnearring INI\subseteq N is a right ideal if II is a normal additive subgroup and for all iIi\in I, nNn\in N, inIi*n\in I. A subnearring INI\subseteq N is a left ideal if II is a normal additive subgroup and for all iIi\in I, n,mNn,m\in N, nmn(m+i)In*m-n*(m+i)\in I. A subnearring that is a left and right ideal is an ideal. As we expect, ideals correspond to the kernels of nearring homomorphisms.

We write NN^{*} for N{0}N\setminus\{0\}. If (N,)(N^{*},*) is a group, then (N,+,)(N,+,*) is a nearfield, generalizing fields by having a single distributive law. Nearfields play an important role in geometries, nearring and nearfield theory are well discussed in [10, 11, 17, 19]. Nearfield were first described by Dickson using a process that is now called a Dickson nearfield. All but 7 finite nearfields are Dickson.

The set D(N)={nN|n(a+b)=na+nba,bN}D(N)=\{n\in N|n(a+b)=na+nb\,\forall a,b\in N\} of distributive elements of NN is the core investigation of this paper.

Elements a,bNa,b\in N are called equivalent multipliers if xa=xbx*a=x*b for all xNx\in N. We write aba\cong b, an equivalence relation. A nearring (N,+,)(N,+,*) is called planar if

  • The equivalent multiplier equivalence relation has at least 3 classes

  • For every a,b,cNa,b,c\in N, a≇ba\not\cong b, the equation xa=xb+cx*a=x*b+c has a unique solution.

Every field other than 2\mathbb{Z}_{2} is a planar nearring. A finite nearfield (except 2\mathbb{Z}_{2}) is always planar, there are known to be nonplanar infinite nearfields [19, page 46]. Planar nearrings are zero symmetric.

Planar nearrings can be described by fixed point free automorphism groups of groups. Let (N,+)(N,+) be a group. Then we say that some nonidentity ϕAut(N)\phi\in Aut(N) is fixed point free if for all nNn\in N, nϕ=nn\phi=n iff n=0n=0. A group of automorphisms is fixed point free if all nonidentity elements are fixed point free. Now let ΦAut(N)\Phi\leq Aut(N) be a group of fixed point free automorphisms of NN acting from the right such that id+ϕ-id+\phi is bijective for all non identity ϕΦ\phi\in\Phi. We write aΦa\Phi for the orbit containing aa. Let RNR\subseteq N be a set of orbit representatives, MRM\subseteq R a set. Every element aNa\in N can be written uniquely as raϕar_{a}\phi_{a} for some raRr_{a}\in R, ϕaΦ\phi_{a}\in\Phi. We define a multiplication by:

ab={0rbMra(ϕaϕb)=aϕbrbMa*b=\begin{cases}0&r_{b}\in M\\ r_{a}(\phi_{a}\phi_{b})=a\phi_{b}\,&r_{b}\not\in M\end{cases} (1)

Then (N,+,)(N,+,*) is a planar (right) nearring. All planar nearrings can be so derived[10].

Let aΦ=aΦ{0}a\Phi^{*}=a\Phi\cup\{0\}. We call MM the set of zero multipliers. If aMa\in M then we call aΦa\Phi a zero multiplier orbit, and we will call all elements of this orbit zero multipliers. Then nNn\in N^{*} is a zero multiplier iff nMΦn\in M\Phi^{*} iff an=0a*n=0 for all aNa\in N. Note that ab=0a*b=0 iff a=0a=0, b=0b=0 or rbMr_{b}\in M. In particular, nb=0n*b=0 for all nn iff b=0b=0 or rbMr_{b}\in M. The elements rRMr\in R\setminus M are right identities, xr=xx*r=x for all nNn\in N. A planar nearring has a left identity iff it has an identity iff it has exactly one nontrivial Φ\Phi orbit iff it is a planar nearfield. We use Z(Φ)Z(\Phi) to denote the centre of the group Φ\Phi and remark that we will use the British spelling throughout this paper. The distributive elements of a nearfield are called the kern of the nearfield and contains the multiplicative centre.

There are related nearrings. The trivial nearring on any group (N,+)(N,+) with multiplication ab=0a*b=0 would correspond to \cong having one equivalence class. The Malone trivial nearrings [8] correspond to Φ\Phi having order 1, so R=N{0}R=N\setminus\{0\}, ab=0a*b=0 if bMb\in M and ab=aa*b=a otherwise. It is interesting to note that the complemented Malone nearrings in [8] are a generalization of planar nearrings with Φ\Phi of order 2, allowing fixed points (i.e. elements of additive order 2) that are zero multipliers.

In the next section we will look at two special constructions for planar nearrings. Then in the following section we look at the distributive elements of a planar nearring. With that knowledge, we will determine the generalized centre of a planar nearring.

3 (Near) Vector Spaces

In this section we construct two families of example planar nearrings, which will prove to be useful in the rest of the paper.

Let VV be a vector space over a division ring DD of order at least 3 and ϕ:VD\phi:V\rightarrow D a vector space epimorphism which acts from the right. Define a multiplication :V×VV*:V\times V\rightarrow V by ab=a(bϕ)a*b=a(b\phi). By [2, theorem 4.1] and [21, theorem 5.2.1] this is a planar ring and all planar rings have this form. Using the terminology above, Φ\Phi is isomorphic to the nonzero elements of DD under multiplication, RM=ϕ1(1)R\setminus M=\phi^{-1}(1) and the elements of MM can be chosen arbitrarily from the orbits that lie completely within kerϕ\ker\phi. Let v1Rv_{1}\in R be arbitrary. If VV is a finite dimensional vector space, we can choose a new basis v1,,vnVv_{1},\dots,v_{n}\in V such that v2,,vnkerϕv_{2},\dots,v_{n}\in\ker\phi and for all x=(x1,,xn)Vx=(x_{1},\dots,x_{n})\in V, xϕ=x1x\phi=x_{1}.

We can generalize this construction to nearvector spaces111We note in passing the existance of another, similar, definition of a nearvector space used by Karzel and colleagues[12, 13] in which the right nearfield scalars operate from the left, giving significantly different properties.. We begin with a brief overview and some definitions of nearvector spaces. See [5] for further details.

Definition 3.1.

A pair (V,A)(V,A) is called a nearvector space if:

  1. 1.

    (V,+)(V,+) is a group and AA is a set of endomorphisms of VV, which act from the right;

  2. 2.

    AA contains the endomorphisms 0, id and -id;

  3. 3.

    A=A{0}A^{*}=A\setminus\{0\} is a subgroup of the group Aut(V)(V);

  4. 4.

    AA acts fixed point freely on VV;

  5. 5.

    the quasi-kernel {xV|α,βA,γA:xα+xβ=xγ}\{x\in V\,|\,\forall\alpha,\beta\in A,\exists\gamma\in A:x\alpha+x\beta=x\gamma\} generates VV as a group.

We sometimes refer to VV as a nearvector space over AA. We write Q(V)Q(V) for the quasi-kernel of VV. The elements of VV are called vectors and the members of AA scalars and it turns out that AA is a nearfield. The action of AA on VV is called scalar multiplication. Note that (V,+)(V,+) is an abelian group. Also, the dimension of the nearvector space, dim(V)\dim(V), is uniquely determined by the cardinality of an independent generating set for Q(V)Q(V).

In [18, theorem 3.4] we find a characterization of finite dimensional nearvector spaces, see also [5, theorem 4.6].

Theorem 3.2.

Let VV be a group and let A:=D{0}A:=D\,\cup\,\{0\}, where DD is a fixed point free group of automorphisms of VV. Then (V,A)(V,A) is a finite dimensional nearvector space if and only if there exists a finite number of nearfields, F1,F2,,FnF_{1},F_{2},\ldots,F_{n}, semigroup isomorphisms ψi:AFi\psi_{i}:A\rightarrow F_{i} and a group isomorphism Φ:VF1F2Fn\Phi:V\rightarrow F_{1}\oplus F_{2}\oplus\dots\oplus F_{n} such that if

Φ(v)=(x1,x2,,xn),(xiFi)\Phi(v)=(x_{1},x_{2},\dots,x_{n}),\,\,\,(x_{i}\in F_{i})

then

Φ(vα)=(x1(αψ1),x2(αψ2),,xn(αψn)),\Phi(v\alpha)=(x_{1}(\alpha\psi_{1}),x_{2}(\alpha\psi_{2}),\dots,x_{n}(\alpha\psi_{n})),

for all vVv\in V and αA\alpha\in A.

In [5, 4.13 ff] we find the following. A nearvector space is regular if all the ψi\psi_{i} are identical (up to nearfield automorphisms). Every nearvector space has a unique maximal decomposition V=V1V2V=V_{1}\oplus V_{2}\oplus\dots into regular sub-nearvector spaces ViV_{i}. Then for all nonzero uQ(V)u\in Q(V), there is precisely one ii auch that uViu\in V_{i}. Note that a regular vector space over a field is a vector space.

We can construct a planar nearring from a near vector space along the lines used above for vector spaces. Let VV be a nearvector space over a nearfield FF of order at least 3. Let ϕ:VF\phi:V\rightarrow F be a nearvector space epimorphism and define ab=a(bϕ)a*b=a(b\phi). The right identities RMR\setminus M are ϕ1(1)\phi^{-1}(1) and the representatives in MM can be chosen arbitrarily.

Example 3.3.

Let VV be the two dimensional nearvector space over F=5F=\mathbb{Z}_{5} with ψ1\psi_{1} the identity and ψ2=(2 3)\psi_{2}=(2\,3) the automorphism of FF^{*} exchanging 22 and 33, equivalently xψ2=x3x\psi_{2}=x^{3}. Note that Q(V)=F×{0}{0}×FQ(V)=F\times\{0\}\cup\{0\}\times F. Taking ϕ(v1,v2)=v1\phi(v_{1},v_{2})=v_{1}, we obtain a planar nearring (V,+,)(V,+,*). Then (v1,v2)D(V)(v_{1},v_{2})\in D(V) iff v1x+v1y=v1(x+y)v_{1}x+v_{1}y=v_{1}(x+y) and v2(xψ2)+v2(yψ2)=v2((x+y)ψ2)v_{2}(x\psi_{2})+v_{2}(y\psi_{2})=v_{2}((x+y)\psi_{2}) for all x,y,Fx,y,\in F. The first equation always holds, but the second equation can be seen to fail for x=y=1x=y=1 unless v2=0v_{2}=0. Thus we see that D(V)=F×{0}D(V)=F\times\{0\}.

Example 3.4.

Let VV be the two dimensional nearvector space over FF, the proper nearfield of order 9 with kern KK of order 3. Let ψ1=ψ2\psi_{1}=\psi_{2} be the identity. Then Q(V)=K×KQ(V)=K\times K. Taking ϕ(v1,v2)=v1\phi(v_{1},v_{2})=v_{1}, we obtain a planar nearring (V,+,)(V,+,*). Then (v1,v2)D(V)(v_{1},v_{2})\in D(V) iff v1x+v1y=v1(x+y)v_{1}x+v_{1}y=v_{1}(x+y) and v2x+v2y=v2(x+y)v_{2}x+v_{2}y=v_{2}(x+y) for all x,y,Fx,y,\in F. These equations hold iff v1v_{1} and v2v_{2} are both in the kern KK of FF, so D(V)=K×KD(V)=K\times K.

Conjecture 3.5.

Let FF be a nearfield with kern KK, let VV be a finite dimensional FF-nearvector space derived planar nearring as above, V=V1VnV=V_{1}\oplus\dots\oplus V_{n} the regular decomposition with Vi=FniV_{i}=F^{n_{i}}. Then D(V)=D(V1)D(Vn)D(V)=D(V_{1})\oplus\dots\oplus D(V_{n}) with D(Vi)=KniD(V_{i})=K^{n_{i}}.

It is worth noting in passing that nearvector spaces and the homogeneous mappings of them to themselves are closely related to questions about nearring matrices over the associated nearfield. Thus we hope that future work here could shed light on the question raised in the final section of [14] as to the inverses of units in matrix nearrings over planar nearfields.

4 The Distributive Elements

In this section we investigate the distributive elements of a planar nearring. Some examples are well known. A finite field is a planar nearfield and thus a planar nearring, with all elements being distributive. In [2] (see §3 above) the structure of planar rings is completely determined, so we know what happens when D(N)=ND(N)=N. The distributive elements of a nearfield are called the kern of the nearfield.

Using Sonata [1] we found all planar nearrings up to order 15 with nontrivial D(N)D(N).

  1. 1.

    The fields of order 3,4,5,7,8,9,11,13.

  2. 2.

    The proper nearfield of order 9 with kern of order 3.

  3. 3.

    The planar ring of order 9.

  4. 4.

    An example of order 9. The additive group is 9\mathbb{Z}_{9}, Φ={1,1}\Phi=\{1,-1\}, R={2,3,5,8}R=\{2,3,5,8\}, M={3}M=\{3\}. The distributive elements are the zero multipliers {0,3,6}\{0,3,6\}.

  5. 5.

    An example of order 15, Φ\Phi of order 2 with generator gg acting on 3×5\mathbb{Z}_{3}\times\mathbb{Z}_{5} as (x,y)g=(x,y)(x,y)*g=(-x,-y), the orbit {0}×5\{0\}\times\mathbb{Z}_{5} zero multipliers. The distributive elements are 3×{0}\mathbb{Z}_{3}\times\{0\}.

We see that 1-3 can be readily explained, but not 4 or 5. One of the goals of this paper is to understand all these examples in terms of general classes.

We first determine some properties of the orbits that contain distributive elements.

Lemma 4.1.

Let dNd\in N. Then dD(N)dΦd\in D(N)\Rightarrow d\Phi^{*} is additively closed.

Proof 4.2.

Suppose dΦd\Phi^{*} is not additively closed, so that there exist some ϕ1,ϕ2\phi_{1},\phi_{2} such that dϕ1+dϕ2dΦd\phi_{1}+d\phi_{2}\not\in d\Phi^{*}. Let r3ϕ3=rdϕ1+rdϕ2r_{3}\phi_{3}=r_{d}\phi_{1}+r_{d}\phi_{2}. Then d(rdϕ1+rdϕ2)=dr3ϕ3=rdϕdϕ3dΦd*(r_{d}\phi_{1}+r_{d}\phi_{2})=d*r_{3}\phi_{3}=r_{d}\phi_{d}\phi_{3}\in d\Phi^{*}. However drdϕ1+drdϕ2=dϕ1+dϕ2dΦd*r_{d}\phi_{1}+d*r_{d}\phi_{2}=d*\phi_{1}+d*\phi_{2}\not\in d\Phi^{*}, so dd is not distributive, a contradiction.

Lemma 4.3.

Let NN be a planar nearring. Let dD(N)d\in D(N) be a non zero multiplier, d=rdϕdd=r_{d}\phi_{d}. Then {ϕΦ|rdϕD(N)}Φ\{\phi\in\Phi|r_{d}\phi\in D(N)\}\leq\Phi is a subgroup, containing Z(Φ)Z(\Phi).

Proof 4.4.

From lemma 4.1 above, we know that dΦd\Phi^{*} is additively closed.

Let a,bNa,b\in N, ϕΦ\phi\in\Phi. Let r=rdr=r_{d}. Then we can show the following.

rϕ(a+b)\displaystyle r\phi*(a+b) =rϕrϕd1rϕd(a+b)\displaystyle=r\phi*r\phi_{d}^{-1}*r\phi_{d}*(a+b) (2)
=rϕrϕd1(rϕda+rϕdb)\displaystyle=r\phi*r\phi_{d}^{-1}*(r\phi_{d}*a+r\phi_{d}*b) (3)
=rϕrϕd1(rϕdra+rϕdrb)\displaystyle=r\phi*r\phi_{d}^{-1}*(r\phi_{d}*r*a+r\phi_{d}*r*b) (4)
=rϕrϕd1rϕd(ra+rb)\displaystyle=r\phi*r\phi_{d}^{-1}*r\phi_{d}*(r*a+r*b) (5)
=rϕ(ra+rb)\displaystyle=r\phi*(r*a+r*b) (6)

We use this in the following calculation. We know that rdΦr_{d}\Phi^{*} is additively closed and that r=rdr=r_{d} is a left multiplicative identity in rdΦr_{d}\Phi.

rϕd1a+rϕd1b\displaystyle r\phi_{d}^{-1}*a+r\phi_{d}^{-1}*b =r(rϕd1a+rϕd1b)\displaystyle=r*(r\phi_{d}^{-1}*a+r\phi_{d}^{-1}*b) (7)
=rϕd1rϕd(rϕd1a+rϕd1b)\displaystyle=r\phi_{d}^{-1}*r\phi_{d}*(r\phi_{d}^{-1}*a+r\phi_{d}^{-1}*b) (8)
=rϕd1(rϕdrϕd1a+rϕdrϕd1b)\displaystyle=r\phi_{d}^{-1}*(r\phi_{d}*r\phi_{d}^{-1}*a+r\phi_{d}*r\phi_{d}^{-1}*b) (9)
=rϕd1(ra+rb)\displaystyle=r\phi_{d}^{-1}*(r*a+r*b) (10)
=rϕd1(a+b)\displaystyle=r\phi_{d}^{-1}*(a+b) (11)

Thus the multiplicative inverse of dd in dΦd\Phi is in D(N)dΦD(N)\cap d\Phi. By standard arguments, D(N)dΦD(N)\cap d\Phi is multiplicatively closed, thus a group, so{ϕΦ|rdϕD(N)}Φ\{\phi\in\Phi|r_{d}\phi\in D(N)\}\leq\Phi is a subgroup of Φ\Phi.

Now we know that rd(a+b)=rda+rdbr_{d}*(a+b)=r_{d}*a+r_{d}*b. Let ϕZ(Φ)\phi\in Z(\Phi). Then let a,bNa,b\in N,

rdϕ(a+b)\displaystyle r_{d}\phi*(a+b) =rdϕϕa+b\displaystyle=r_{d}\phi\phi_{a+b} (12)
=rdϕa+bϕ\displaystyle=r_{d}\phi_{a+b}\phi (13)
=rd(a+b)ϕ\displaystyle=r_{d}(a+b)\phi (14)
=rdϕa+rdϕb\displaystyle=r_{d}\phi*a+r_{d}\phi*b (15)

so Z(Φ){ϕΦ|rdϕD(N)}Z(\Phi)\leq\{\phi\in\Phi|r_{d}\phi\in D(N)\} and we are done.

Lemma 4.5.

Let dNd\in N be a non zero multiplier. Then dD(N)dΦd\in D(N)\Rightarrow d\Phi^{*} is a planar nearfield.

Proof 4.6.

From lemma 4.1 above, we know that dΦd\Phi^{*} is additively closed. Thus dΦd\Phi^{*} is additively and multiplicatively closed, forming a planar subnearring. We note that there is precisely one orbit on this nearring, so the planar nearring must be a nearfield with rdr_{d} the multiplicative identity.

Let n,cD(N)n,c\in D(N) be arbitrary, then there exists a unique xNx\in N such that xxn=cx-x*n=c by the planarity of NN, but this might not be in dΦd\Phi^{*}. However rdxrdxn=rd(xxn)=rdc=cr_{d}*x-r_{d}*x*n=r_{d}*(x-x*n)=r_{d}*c=c so rdxr_{d}*x is also a solution to the equation. This solution is unique so x=rdxdΦx=r_{d}*x\in d\Phi^{*} and dΦd\Phi^{*} is a planar nearfield.

Similarly we know that, even in the case that there are distributive elements that are zero multipliers, additive closure of the orbit dΦd\Phi^{*} allows us to define a multiplication dϕ1dϕ2=d(ϕ1ϕ2)d\phi_{1}\circ d\phi_{2}=d(\phi_{1}\phi_{2}) that gives us (dΦ,+,)(d\Phi^{*},+,\circ) a planar nearfield.

Thus we know a lot more about the forms of Φ\Phi that can emerge, as not all fixed point free automorphism groups arise as the multiplicative group of a nearfield.

Lemma 4.7.

Let NN be a planar nearring with nontrivial distributive elements. Then for every mMΦm\in M\Phi^{*}, aNMΦa\in N\setminus M\Phi^{*}, ϕm+a=ϕa\phi_{m+a}=\phi_{a}.

Proof 4.8.

We proceed by calculation. Let 0dD(N)0\neq d\in D(N).

rdϕdϕm+a\displaystyle r_{d}\phi_{d}\phi_{m+a} =d(m+a)=dm+da\displaystyle=d*(m+a)=d*m+d*a (16)
=0+da=da=rdϕdϕa\displaystyle=0+d*a=d*a=r_{d}\phi_{d}\phi_{a} (17)
ϕdϕm+a\displaystyle\Rightarrow\phi_{d}\phi_{m+a} =ϕdϕa\displaystyle=\phi_{d}\phi_{a} (18)
ϕm+a\displaystyle\Rightarrow\phi_{m+a} =ϕa\displaystyle=\phi_{a} (19)

By symmetry, ϕa+m=ϕa\phi_{a+m}=\phi_{a} as well.

Lemma 4.9.

Let NN be a planar nearring. If D(N)D(N) is nontrivial, then the zero multipliers form a nearring ideal.

Proof 4.10.

Let 0dD(N)0\neq d\in D(N). Let K=MΦK=M\Phi^{*}, the zero multipliers. The mapping ρ:ndn\rho:n\mapsto d*n is an additive homomorphism. Then nkerρn\in\ker\rho iff dn=0d*n=0 iff nKn\in K so the zero multipliers form an additively normal group.

Let m,nNm,n\in N, kKk\in K. If nn is a zero multiplier, then kn=0k*n=0 so knKk*n\in K. If nn is not a zero multiplier, then kn=rkϕkϕnKk*n=r_{k}\phi_{k}\phi_{n}\in K, so KK is a right ideal. If n+kKn+k\in K, then since KK is a subgroup, nKn\in K so mnm(n+k)=00Km*n-m*(n+k)=0-0\in K. If n+kKn+k\not\in K, then remember that by the lemma above, ϕn+k=ϕn\phi_{n+k}=\phi_{n}. Then

mnm(n+k)\displaystyle m*n-m*(n+k) =mnrmϕmϕn+k\displaystyle=m*n-r_{m}\phi_{m}\phi_{n+k} (20)
=mnrmϕmϕn\displaystyle=m*n-r_{m}\phi_{m}\phi_{n} (21)
=mnmn=0K\displaystyle=m*n-m*n=0\in K (22)

so we have a left ideal and thus an ideal.

Note that, in general, the mapping ρ\rho is not a nearring homomorphism, unless ϕd\phi_{d} is the identity. This can only be guaranteed to be the case when D(N)D(N) contains non zero multipliers, by lemma 4.3.

This implies that (N,+)(N,+) is an extension of (ρ(N),+)(\rho(N),+) by (K,+)(K,+). By lemma 4.5 and the comments afterwards, we know that (ρ(N),+)(\rho(N),+) is the additive group of a nearfield, thus abelian and, in the finite case, elementary abelian. When all distributive elements are zero multipliers, we do not necessarily have that the extension splits. If we have a non zero multiplier distributive element, then we get a clear result.

Theorem 1.

Let NN be a planar nearring with automorphism group Φ\Phi. Let dD(N)d\in D(N) be a non zero multiplier. Then there is a subnearfield FNF\leq N with FΦF^{*}\cong\Phi and an additive group KK with Φ\Phi a group of fixed point free automorphisms, such that NKFN\cong K\rtimes F as an additive group and

(a,b)(c,d)\displaystyle(a,b)*(c,d) ={0d=0(aϕd,bϕd)otherwise\displaystyle=\begin{cases}0&d=0\\ (a\phi_{d},b\phi_{d})&\mbox{otherwise}\end{cases} (23)

such that (N,+,)(KF,+,)(N,+,*)\cong(K\rtimes F,+,*). The zero multipliers form an ideal K×{0}K\times\{0\}.

Proof 4.11.

By the previous lemma, we know that (N,+)(N,+) is an extension of (K,+)(K,+) by (ρ(N),+)(\rho(N),+). By lemma 4.5 we know that ρ(N)\rho(N) is a nearfield, let F=ρ(N)F=\rho(N). Because (F,+)(F,+) is a subgroup of NN that is fixed by ρ\rho, the extension splits, so (N,+)(K,+)(F,+)(N,+)\cong(K,+)\rtimes(F,+). Let kKk\in K, fFf\in F. Writing kf=f+kfk^{f}=f+k-f we know that for all k1,k2Kk_{1},k_{2}\in K, f1,f2Ff_{1},f_{2}\in F, (k1,f1)+(k2,f2)=(k1+k2f1,f1+f2)(k_{1},f_{1})+(k_{2},f_{2})=(k_{1}+k_{2}^{f_{1}},f_{1}+f_{2}).

We can write each element of NN as (k,f)=(k,0)+(0,f)MΦ+F(k,f)=(k,0)+(0,f)\in M\Phi^{*}+F so by lemma 4.7 we know that ϕ(k,f)=ϕ(0,f)\phi_{(k,f)}=\phi_{(0,f)}. We write ϕf\phi_{f} for ϕ(0,f)\phi_{(0,f)}. Then we can write the multiplication on KFK\rtimes F as above. The representatives for KFK\rtimes F are {(k,1)|kK}{(m,0)|mM}\{(k,1)|k\in K\}\cup\{(m,0)|m\in M\}. Since the additive groups are isomorphic as Φ\Phi-groups and the representatives are matched by the isomorphism, we know that the resulting planar nearrings are isomorphic.

We see that the example on additive group (9,+)(\mathbb{Z}_{9},+) falls outside this theorem. The distributive elements lie within the zero multipliers and the additive group is not a semidirect product of the zero multipliers with anything.

We can create a family of examples of planar nearrings with D(N)D(N) lying within the zero multipliers, based upon the example on page 49 of [10]. These examples do not split.

Example 4.12.

Let pp be an odd prime, N=p2N=\mathbb{Z}_{p^{2}} the cyclic group of order p2p^{2}. There is a cyclic subgroup Φ\Phi of the multiplicative semigroup of order p1p-1. One of the orbits of this automorphism group is pp2p\mathbb{Z}_{p^{2}}. These are our zero multipliers, this orbit has representative pp2p\in\mathbb{Z}_{p^{2}}. We choose the rest of our representatives to be a coset of pp2p\mathbb{Z}_{p^{2}}. Then the resulting planar nearring has D(N)=pp2D(N)=p\mathbb{Z}_{p^{2}}, all zero multipliers.

We know (e.g. [16]) that the additive group of a finite planar nearring is nilpotent and thus a direct sum of pp-groups. Thus a finite planar nearring is a finite direct sum of planar nearrings of prime power order. Thus by lemma 4.5 at most one of these summands has a non zero multiplier distributive element. If one summand has such an element, then lemma 4.9 indicates that we have a trivial multiplication of all summands other than the one with a non zero multiplier distributive element. We have shown the following.

Corollary 2.

Let NN be a finite planar nearring with nontrivial distributive elements. Let Φ\Phi be the multiplicative group associated to NN, RR the representatives and MM the zero multipliers. Then (N,+)(N,+) is the direct sum of finitely many p-subgroups N=N1NkN=N_{1}\oplus\dots\oplus N_{k}, only one of which has a nontrivial multiplication, so RM{0}{0}Ni{0}{0}R\setminus M\subseteq\{0\}\oplus\dots\oplus\{0\}\oplus N_{i}\oplus\{0\}\oplus\dots\oplus\{0\} for some ii.

In the example of order 15, we have the 3\mathbb{Z}_{3} as the nearfield with automorphism group of order 2 and 5\mathbb{Z}_{5} having the same automorphism group generated by (1 4)(2 3)(1\,4)(2\,3) acting on it. Then N=3×5N=\mathbb{Z}_{3}\times\mathbb{Z}_{5} with {0}×5\{0\}\times\mathbb{Z}_{5} forming the zero multipliers in the nearring.

Thus we have shown all of our small examples of planar nearrings with nontrivial distributive elements fall into larger classes of examples.

5 Some applications

In this section we look at the way that these results can be contextualized in relation to other similar results.

One of the strong applications of planar nearrings is in the construction of BIBDs. The blocks of the BIBD are {aΦ+b|0aN,bN}\{a\Phi^{*}+b|0\neq a\in N,b\in N\} and the basic blocks are {aΦ|0aN}\{a\Phi^{*}|0\neq a\in N\}. The following result shows that the construction used in [9] is the only way that all basic blocks can be subgroups.

Lemma 5.1.

Let NN be a finite abelian planar nearring with more than one orbit, in which all orbits aΦa\Phi^{*} are additively closed. Then NN is a vector space over a subfield of NN.

Proof 5.2.

Let F=aΦF=a\Phi^{*} for some non zero multiplier aa. FF is additively and multiplicatively closed, thus a planar subnearring. FF contains only one orbit of Φ\Phi, so it has an identity and is thus a nearfield. Thus FF is prime power order. If FF is odd, then Φ\Phi is even order, thus 1Φ-1\in\Phi. If FF is even, then 1=1Φ-1=1\in\Phi.

We see that Φ\Phi^{*} acting on (N,+)(N,+) satisfies the first four conditions for being a nearvector space. By the additive closedness of each nΦn\Phi^{*}, for each α,βΦ\alpha,\beta\in\Phi^{*}, nα+nβnΦn\alpha+n\beta\in n\Phi^{*} so there exists some γΦ\gamma\in\Phi^{*} such that nα+nβ=nγn\alpha+n\beta=n\gamma. Thus Q(N)=NQ(N)=N and NN is a near vector space over Φ\Phi^{*}. We see that FF must then be the nearfield from van der Walt’s result above, (Φ,)(F,)(\Phi^{*},*)\cong(F,*). By [5, Satz 5.5] we know that Q(N)=NQ(N)=N implies that NN is a vector space and that FF is actually a field.

We note in passing that if a planar nearring has non trivial distributive elements, then we know that the corresponding orbits give additively closed basic blocks. Thus by [10, Thm 7.17] we will obtain a statistical, but not a geometric BIBD.

We also obtain Aichinger’s result as a corollary.

Corollary 1.

Let NN be a planar ring. Then NN is a vector space over a field FF with Φ=F\Phi=F^{*}.

Proof 5.3.

All elements of NN are distributive, so we know by lemma 4.1 that every orbit is additively closed. Select some non zero multiplier dNd\in N and we see that dNd*N is a distributive nearfield, that is, a field. All orbits are additively closed, so we obtain a nearvector space with Q(N)=NQ(N)=N, so by the same argument as above, we know that NN is a vector space.

6 The Generalized Centre

Now that we know some things about the distributive elements of a planar nearring, we can say some things about the generalized centre.

Let (N,+,)(N,+,*) be a 0-symmetric nearring. The generalized centre of NN is GC(N)={nN|nd=dndD(N)}GC(N)=\{n\in N|nd=dn\;\forall d\in D(N)\} [3, 7].

The generalized centre was introduced because the centre of a nearring is not always well behaved. For instance, it is not always a subnearring, while the generalized centre is. If NN is a ring, then GC(N)GC(N) is the usual centre of NN.

The generalizes centre of a planar nearfield FF is the set of elements that commute with the kern. In the finite case, the kern is the multiplicative centre and thus GC(F)=FGC(F)=F. In the infinite case, when the kern is distinct from the multiplicative centre, we know only that GC(F)GC(F) contains the multiplicative centre.

Theorem 1.

Let NN be a planar nearring. Then GC(N)GC(N) is one of four cases:

  1. 1.

    If D(N)D(N) intersects only zero multiplier orbits, then GC(N)GC(N) is the zero multipliers, an ideal.

  2. 2.

    If D(N)D(N) intersects more than one orbit of Φ\Phi and at least one of them is not a zero multiplier, then GC(N)={0}GC(N)=\{0\}.

  3. 3.

    If D(N)D(N) intersects exactly one orbit of Φ\Phi, which is not a zero multiplier, let it be aΦa\Phi. Then aZ(Φ)GC(N)aΦaZ(\Phi)^{*}\leq GC(N)\leq a\Phi^{*}.

  4. 4.

    If D(N)={0}D(N)=\{0\}, then GC(N)=NGC(N)=N.

Proof 6.1.

We proceed by cases.

Case 1: Suppose D(N)D(N) intersects only zero multiplier orbits. Then for all dD(N)d\in D(N), for all nNn\in N, nd=0nd=0. Thus if rnMr_{n}\in M, dn=0dn=0 and thus nGC(N)n\in GC(N). So GC(N)={aΦ|aM}GC(N)=\cup\{a\Phi^{*}|a\in M\}, the zero multipliers, which we know from lemma 4.9 to be an ideal.

Case 2: Suppose D(N)D(N) intersects two orbits nontrivially, one of them is a non zero multiplier orbit. Let a,bD(N)a,b\in D(N) be in distinct orbits, aa not a zero multiplier. Then a nonzero cGC(N)c\in GC(N) implies that ca=acca=ac, so ra=rcr_{a}=r_{c} and cc is not a zero multiplier. Thus cb=bccb=bc implies that bb is also not a zero multiplier, so rc=rbr_{c}=r_{b}, but rarbr_{a}\neq r_{b} so we have a contradiction, so GC(N)GC(N) is trivial.

Case 3: Let cGC(N)c\in GC(N), so cd=dccd=dc for all dD(N)d\in D(N). Then rc=rdr_{c}=r_{d} so caΦc\in a\Phi^{*}, giving us the upper bound. We know that F:=aΦF:=a\Phi^{*} is a nearfield, so D(N)=KD(N)=K is the kern of FF. If KFK^{*}\mathrel{\unlhd}F^{*} as multiplicative groups, then K=Z(F)K=Z(F) by [4], so GC(N)=F=aΦGC(N)=F=a\Phi^{*} showing that this bound can be achieved, for instance in the finite case. Otherwise ϕcZ(Φ)\phi_{c}\in Z(\Phi) gives us the lower bound.

Case 4: If D(N)D(N) is trivial, then by zerosymmetry, GC(N)=NGC(N)=N.

We can break this down depending upon the properties of the additive group.

Corollary 2.

Let NN be a planar nearring with nonabelian additive group. Then GC(N)GC(N) is one of four cases:

  1. 1.

    if D(N)D(N) intersects only zero multiplier orbits, then GC(N)GC(N) is the zero multipliers, an ideal.

  2. 2.

    If D(N)D(N) intersects more than one orbit of Φ\Phi and at least one of them is not a zero multiplier, then GC(N)={0}GC(N)=\{0\}.

  3. 3.

    If D(N)D(N) intersects exactly one orbit of Φ\Phi, which is not a zero multiplier, let it be aΦa\Phi, then GC(N)=D(N)=aΦGC(N)=D(N)=a\Phi^{*}.

  4. 4.

    If D(N)D(N) intersects no nonzero orbit of Φ\Phi, then GC(N)=NGC(N)=N.

Proof 6.2.

Only the third case is different to the above. If NN is additively nonabelian, then we know that the fixed point free automorphism group is of odd order, thus cyclic, see e.g. [16]. Thus Z(Φ)=ΦZ(\Phi)=\Phi, so D(N)D(N) is all of one orbit with zero. Because the multiplication within this orbit is commutative, the generalized centre is all of the orbit and we are done.

If the additive group is abelian, then many interesting and strange things can happen with skew fields. However in the finite case we know more.

Corollary 3.

Let NN be a finite planar nearring with abelian additive group. Then GC(N)GC(N) is one of four cases:

  1. 1.

    If D(N)D(N) intersects only zero multiplier orbits, then GC(N)GC(N) is the zero multipliers, an ideal.

  2. 2.

    If D(N)D(N) intersects more than one orbit of Φ\Phi and at least one of them is not a zero multiplier, then GC(N)={0}GC(N)=\{0\}.

  3. 3.

    If D(N)D(N) intersects exactly one orbit of Φ\Phi, which is not a zero multiplier, let it be aΦa\Phi. Then GC(N)=aΦD(N)GC(N)=a\Phi^{*}\geq D(N).

  4. 4.

    If D(N)D(N) intersects no nonzero orbit of Φ\Phi, then GC(N)=NGC(N)=N.

Proof 6.3.

Only the third case is different to the above. We note that abelian addition implies that the distributor D(N)D(N) is additively closed. Thus D(N)D(N) is a planar ring. By [2] we know that a planar ring is derived from a vector space over a field, where the field multiplication is isomorphic to the fixed point free automorphisms of the planar ring. We know that D(N)D(N) lies within one orbit which is a nearfield, so (D(N),)(D(N)^{*},\cdot) is a cyclic subgroup of Φ\Phi and thus precisely the centre of Φ\Phi. Thus all elements of the orbit containing the distributor are in the generalized centre.

7 Conclusion

In this paper we have investigated the distributive elements in a planar nearring. We have been able to show that if there are nontrivial distributive elements then the additive group is an extension of an abelian subgroup by the zero multipliers. This additive group is the additive group of a nearfield, so elementary abelian in the finite case. If the distributive elements include non zero multipliers, then the extension splits and we obtain a clear structure.

As a result, we are able to re-prove Aichinger’s theorem on planar rings as a corollary, as well as Clay’s results on BIBDs with additively closed basic blocks. It is unclear whether lemma 5.1 can be extended to the infinite case. Applying these results to the question of the generalized centre, we are able to obtain a clear set of cases and to describe the generalized centre.

It would be valuable to know what sort of other examples can occur with the distributive elements all lying within the zero multipliers, in order to complete the classification of structures.

It would be of value to calculate D(N)D(N) explicitly in theorem 1. It is easy to see that D(N)D(N) is a direct product of some subset EKE\subseteq K and the kernel of FF. The question is how to calculate which parts of KK have a(ϕb+c)=aϕb+aϕca(\phi_{b+c})=a\phi_{b}+a\phi_{c} where the first addition is in FF while the second is in KK. Note that the orbits will be additively closed, giving us nearfields. This might be another nearvector space construction.

As we have been able to determine the generalized centre of planar nearrings, we can now look forward to describing the generalized centre of more complex classes of nearrings.

8 Acknowledgements

Research supported by SFB Project F5004 of the Austrian Science Foundation, FWF. I would like to thank my colleagues Günter Pilz and Wen-Fong Ke for some insightful questions in the early development of this paper.

References

  • [1] E. Aichinger, F. Binder, J. Ecker, P. Mayr, and C. Nöbauer. SONATA - system of near-rings and their applications, GAP package, Version 2.6, 2012. (http://www.algebra.uni-linz.ac.at/Sonata/).
  • [2] E. Aichinger. Planar rings. Results Math., 30(1-2):10–15, 1996.
  • [3] E. Aichinger and M. Farag. On when the multiplicative center of a near-ring is a subnear-ring. Aequationes Math., 68(1-2):46–59, 2004.
  • [4] J. André. Uber eine Beziehung zwischen Zentrum und Kern endlicher Fastkörper. Archiv der Mathematik, 14(1):145–146, 1963.
  • [5] J. André. Lineare Algebra über Fastkörpern. Math. Z., 136:295–313, 1974.
  • [6] B. Bäck, H. Köppl, G. Pilz, and G. Wendt. Einflußverschiedener Parameter auf den Mykotoxingehalt von Winterweizen Versuchsdurchführung mit Hilfe eines neuen statistischen Modells (Influence of various parameters on the mycotoxin content of winter wheat: experimental process with the assistance of a new statistical model). In Proceedings, 63th ALVA-Tagung, Raumberg, Austria, 2008.
  • [7] G. A. Cannon, M. Farag, and L. Kabza. Centers and generalized centers of near-rings. Comm. Algebra, 35(2):443–453, 2007.
  • [8] G. A. Cannon, M. Farag, L. Kabza, and K. M. Neuerburg. Centers and generalized centers of near-rings without identity defined via Malone-like multiplications. Mathematica Pannonica, to appear.
  • [9] J. R. Clay. Generating balanced incomplete block designs from planar near rings. J. Algebra, 22:319–331, 1972.
  • [10] J. R. Clay. Nearrings, Geneses and applications. (Oxford Science Publications. The Clarendon Press, Oxford University Press, New York, 1992).
  • [11] C. Cotti Ferrero and G. Ferrero. Nearrings, Some developments linked to semigroups and groups, volume 4 of Advances in Mathematics. (Kluwer Academic Publishers, Dordrecht, 2002).
  • [12] H. Karzel. Fastvektorräume, unvollständige Fastkörper und ihre abgeleiteten geometrischen Strukturen. Mitt. Math. Sem. Giessen, (166):127–139, 1984.
  • [13] H. Karzel and G. Kist. Determination of all near vector spaces with projective and affine fibrations. J. Geom., 23(2):124–127, 1984.
  • [14] W.-F. Ke, J. H. Meyer, and G. Wendt. Matrix maps over planar near-rings. Proc. Roy. Soc. Edinburgh Sect. A, 140(1):83–99, 2010.
  • [15] W.-F. Ke and G. Pilz. Abstract algebra in statistics. Journal of Algebraic Statistics, pages 6–12, 2010.
  • [16] P. Mayr. Fixed point free automorphism groups. Master’s thesis, Johannes Kepler University Linz, 1998.
  • [17] G. Pilz. Near-rings. The theory and its applications., volume 23 of North-Holland Mathematics Studies. (North-Holland Publishing Co., Amsterdam, second edition, 1983).
  • [18] A. P. J. van der Walt. Matrix near-rings contained in 22-primitive near-rings with minimal subgroups. J. Algebra, 148(2):296–304, 1992.
  • [19] H. Wähling. Theorie der Fastkörper. (Thales-Verlag, Essen, 1987).
  • [20] G. Wendt. Minimal left ideals of near-rings. Acta Math. Hungar., 127(1-2):52–63, 2010.
  • [21] G. Wendt. Planarity in Near-Rings. PhD thesis, Johannes Kepler University Linz, 2004.