• 検索結果がありません。

Turning the construction ofM∞ upside-down for the two-element left-zero band, we exhibit a duality for quasi-regular left-normal bands

N/A
N/A
Protected

Academic year: 2022

シェア "Turning the construction ofM∞ upside-down for the two-element left-zero band, we exhibit a duality for quasi-regular left-normal bands"

Copied!
24
0
0

読み込み中.... (全文を見る)

全文

(1)

Vol. LXVIII, 2(1999), pp. 295–318

REGULARISING NATURAL DUALITIES

B. A. DAVEY and B. J. KNOX

Abstract. Given an algebraMwe may adjoin an isolated zero to form an algebra Msatisfying all identitiesuvtrue inMfor whichuandvcontain the same variables. Drawing on the structure theory of P lonka sums, we show that ifMis a finite, dualisable algebra which is strongly irregular, thenM is also dualisable.

Turning the construction ofM upside-down for the two-element left-zero band, we exhibit a duality for quasi-regular left-normal bands.

1. Introduction

At the present time it is not clear how common algebraic constructions interact with the theory of natural dualities. For example, it is not known in general whether a finite product of dualisable algebras is dualisable. Even the familiar act of passing to a subalgebra may lead to complications – recently, non-dualisable algebras that may be embedded into dualisable algebras have been discovered (see Clark, Davey and Pitkethly [3]). In this paper we consider the general algebraic analogue of adding a zero to a semigroup and investigate when this construction preserves dualisability.

We aim to take a finite, dualisable algebraMand obtain a natural duality for the quasi-variety generated by the algebra M, formed by adding a zero to M.

This has been achieved by Gierz and Romanowska [7] in the case thatM is the two-element distributive lattice, thus giving an explicit natural duality for the variety of distributive bisemilattices. Romanowska and Smith, in [13] and [14], give a more conceptual treatment of the general case and show that a full (not necessarily natural) duality for a strongly irregular varietyVlifts to a full duality for its regularisation. If V = ISPM and its full duality can be realised using a schizophrenic object M (for example, when the duality is natural), then the induced full duality forISPM can also be realised using a schizophrenic object M. While the algebraic personality ofMis exactlyM, it is not clear how the

Received November 13, 1998; revised July 27, 1999.

1980Mathematics Subject Classification(1991Revision). Primary 08C05, 20M30; Secondary 18A40.

Key words and phrases. natural duality, P lonka sum, regularisation, quasi-regularisation, left-normal semigroup.

The second author was supported by an Australian Postgraduate Research Award.

(2)

structured topological personality ofM may be distilled from the schizophrenic objectM described by Romanowska and Smith.

In this paper, we give direct algebraic paths in the vain of the Gierz-Roma- nowska duality from a (not necessarily full nor strong) natural duality for M to one for M. However, one result assumes almost nothing about the algebra M and may lead to a new binary operation in the type of the adjoined-zero algebra.

This operation turns out to be superfluous exactly whenMis strongly irregular, so in Section 3 we work under this assumption, allowing use of the beautiful structure theory of P lonka sums and regularisations of strongly irregular varieties.

Finally, we exhibit a bare hands natural duality for a semigroup obtained by

“shadowing” an element of the two-element left-zero semigroup. This construction (again a type of P lonka sum) relates to quasi-regularisations of strongly irregular varieties as adjoining a zero relates to regularisations.

We firstly review the setting of Davey and Werner for producing dualities. A leisurely introduction may be found in [5] while [2] gives a detailed account.

LetM be a finite algebra and consider a typeG∪H∪R oftotal operation symbols G, partial operation symbols H, and relation symbols R. Let M∼ =hM;GM, HM, RM;τibe a topological structure having the same underlying set asM, where

(a) each g ∈ G is interpreted as a homomorphismgM: Mn → M for some n∈N∪ {0},

(b) eachh∈H is interpreted as a homomorphismhM: dom(hM)→Mwhere dom(hM) is a subalgebra ofMn for some n∈N,

(c) eachr∈Ris interpreted as a subalgebrarM ofMn for somen∈N, (d) τ is the discrete topology.

Whenever (a), (b) and (c) hold, we say that GM ∪HM ∪RM is algebraic over M. Under these conditions, there is a naturally defined dual adjunction between the quasi-variety A := ISPM and the topological quasi-variety X :=

IScP+M

∼ consisting of isomorphic copies of topologically closed substructures of non-trivial powers of M

∼. For each A∈ A the homset D(A) := A(A,M) (that is, the set of homomorphisms A →M) is a closed substructure ofM

A and for eachX ∈X the homset E(X) := X(X,M

∼) (that is, the set of continuous maps X →M that preserve each total operation, partial operation and relation symbol in G∪H∪R) forms a subalgebra ofMX. It follows that the contravariant hom- functorsA(−,M) :A→SandX(−,M

∼) :X→S, whereSis the category of sets, lift to contravariant functorsD:A→XandE:X→A. For eachA∈A, define theevaluation mapeA:A→ED(A) by

eA(a)(x) :=x(a)

for each a∈ A and each x∈D(A). It may be shown thateA is an embedding for each A ∈ A, as is the similarly defined εX:X → DE(X) for each X ∈ X.

(3)

A simple calculation shows that e: idA → ED and ε: idX → DE are natural transformations.

If, for an algebraA∈A, the embeddingeAis an isomorphism we say thatM yields aduality onA. In caseeA is an isomorphism for all A∈A, we say that∼ M∼ dualises Mor thatGM ∪HM∪RM is adualising structure forM. Here, the natural dual adjunction between A and X is actually a dual representation.

If there is some choice ofGM ∪HM ∪RM such that M

∼ dualisesM, we say that Misdualisable.

One of the aims of natural duality theory is to take the category theoretic and topological conditions for M

∼ to dualise M and distill them into (finitary) algebraic conditions. The following three results follow this programme and will be the foundation of the duality results in this paper. The reader is referred to [2]

for the proofs.

Theorem 1.1(First Duality Theorem). M

∼ yields a duality onA∈A=ISPM if and only if every morphismα:D(A)→M

∼ extends to an A-ary term function t:MA→M of M.

Theorem 1.2(Duality Compactness Theorem). IfM

∼ is of finite type (that is, G∪H ∪R is finite) and yields a duality on each finite algebra A ∈A, then M dualisesM. ∼

Theorem 1.3 (IC Duality Theorem). SupposeG∪H ∪R is finite. ThenM dualisesM provided the followinginterpolation condition is satisfied: ∼

(IC) For each n ∈ N and each substructure X of M

n, every morphism α:X→M

∼ extends to ann-ary term function t:Mn →M of M.

2. Algebras with an Adjoined Isolated Zero

An algebraA of typeF is said to have azeroelement, ∞ ∈A, if there are no nullary operation symbols in F and for every fundamental operation symbol f ∈F (f n-ary) we havefA(x1, . . . , xn) =∞wheneverxi=∞for somei. If, in addition,A:=hA\{∞};Fiis a subalgebra ofA, we call∞anisolatedzero.

Clearly, if there is anf ∈F with arity greater than 1, then an algebra of type F may have at most one zero.

For the remainder of this section we will work in a fixed type F having no nullary operation symbols. Given an algebra, we may adjoin an isolated zero via the following construction.

Definition 1. LetM =hM;Fi be an algebra with∞∈/ M. Define M to be the algebra with universe M =M∪{∞}˙ and fundamental operations given

(4)

by:

fM(x1, . . . , xn) =

fM(x1, . . . , xn), ifx1, . . . , xn∈M,

∞ otherwise,

wheref ∈F isn-ary andx1, . . . , xn ∈M.

We will be considering the situation whereMin the above definition is a finite algebra which is dualised by some structureM

∼ =hM;GM, HM, RM;τi. We would like to dualiseM by some simple modification ofGM ∪HM∪RM.

A result may be obtained using Theorem 3.1 of [6] (see also Theorem 7.7.2 in [2]) in the case where Mhas a one-element subalgebra {a} and a zero bwith a6= b. We may embed M into M2 via the map sending each x∈ M to (x, a) and∞to (b, b), henceM∈ISPM. Also, the mapM →M sending∞toband fixingM is a retraction.

Proposition 2.1. Suppose that

M∼ =hM;GM, HM, RM;τi

dualises M where M has a one-element subalgebra {a} and a zero b with a 6=b.

Then

M∼=hM; End(M), GM∪HM, RM;τi dualisesM.

Returning to the general case, assuming nothing about the algebraMwill lead us to a stronger assumption on M

∼, namely that it satisfies (IC). As a further detour, we will in this section be producing a duality for the algebra

M:=hM;FM∪ {∗}i, where the binary operation∗:M2→M given by

x∗y:=

x ifx, y∈M,

∞ otherwise,

has been added to M as a new fundamental operation. The following Lemma indicates whenM is term equivalent toM.

Lemma 2.2. The binary operation∗ is a term function ofM if and only if M has a left-zero term, that is, a binary term t involving v1 and v2 such that M satisfies the identity

t(v1, v2)≈v1.

Proof. Suppose∗ is a term function ofM, that is,∗ =tM for some binary termt. To see thattinvolves the variablev1, let (x, y),(z, y)∈M2withx, y∈M andz=∞. Then

tM(x, y) =x∗y=x6=∞=z∗y=tM(z, y).

(5)

Similarly, we see that tinvolves v2 by letting (x, y),(x, z)∈M2 withx, y∈M andz=∞and observing

tM(x, y) =x∗y=x6=∞=x∗z=tM(x, z).

Now, for allx, y∈M,

tM(x, y) =tMM2(x, y) =x∗y=x,

asM is a subalgebra ofM, showing that Msatisfies the identityt(v1, v2)≈v1. Conversely, suppose that M has a left-zero term t. We observe that t must contain a fundamental operation symbol of arity at least two, hence by an easy induction on the complexity oft,

tM(x, y) =

tM(x, y) =x ifx, y∈M,

∞ otherwise,

for allx, y∈M, therefore∗=tM.

Note that, regardless of whether or not∗ is artificially introduced, it is imme- diately seen to be a homomorphism M2 →M. Also, for all x, y, z∈M we have

x∗x=x

(x∗y)∗z=x∗(y∗z) x∗y∗z=x∗z∗y.

That is,hM;∗iis aleft normalidempotent semigroup. From the above identi- ties we may obtain theentropiclaw

x∗y∗w∗z=x∗w∗y∗z, so it follows that∗is a homomorphism (M)2→M.

Definition 2. Given a setGM ∪HM∪RM of operations, partial operations and relations algebraic overM, there is a natural way to lift this structure toM via a construction similar to that used in Definition 1.

For eachn-ary g∈G, withn >1, definegM: (M)n →M by:

gM(x1, . . . , xn) =

gM(x1, . . . , xn), ifx1, . . . , xn∈M,

∞ otherwise,

for allx1, . . . , xn∈M. Ifg∈Gis nullary, then we definegM :=gM.

(6)

For eachn-aryh∈H, let dom(hM) = dom(hM) ˙∪{∞}where∞is the constant n-tuple (∞, . . . ,∞) and definehM: dom(hM)→M by:

hM(x1, . . . , xn) =

hM(x1, . . . , xn), ifx1, . . . , xn∈dom(hM),

∞ if (x1, . . . , xn) =∞, for allx1, . . . , xn∈dom(hM).

Finally, for eachn-ary relation symbolr∈R, we let rM =rM∪{∞}˙ .

If GM ∪HM ∪RM is algebraic overM, it follows that the resulting (partial) operations and relationsGM∪HM∪RM arising from Definition 2.3 are alge- braic overM. Further, since our constructions are compatible with∗, we have GM∪HM∪RM algebraic over the extensionM.

Theorem 2.3. LetMbe a finite algebra and supposeGM∪HM∪RM satisfies (IC). Then

M∼:=hM;GM∪ {∗} ∪ {∞}, HM, RM∪ {M};τi satisfies(IC)with respect toM, hence ifGM∪HM∪RM is finite, M

dualises M.

Proof. Letn∈Nand letXbe a (closed) substructure of (M

)nand letλ:X→ M∼ be a morphism. Our goal is to extendλto a term function (M)n →M.

In the first case, ifλ(x) =∞for allx∈X, we must haveX∩Mn=∅sinceλ preserves the unary relationM. That is, for everyx∈X, there is ani∈ {1, . . . , n} such thatxi=∞. It is then easy to see that for allx∈X,x1∗· · ·∗xn=∞=λ(x), showing then-ary term function (v1∗ · · · ∗vn)M extends λ.

Ifλis not the constant map onto{∞}, the set N :={x∈X|λ(x)6=∞}

is non-empty. We define

xN :=x1∗ · · · ∗xl

wherex1, . . . , xlis some fixed sequence of the elements ofN. Sinceλpreserves∗, we must havexN ∈N, for otherwise

λ(xN) =∞ ⇒λ(x1)∗ · · · ∗λ(xl) =∞ ⇒λ(xj) =∞for some xj ∈N , a contradiction. Also, note thatx∈N if and only ifxN ∗x=xN.

(7)

Let

Z :={i∈ {1, . . . , n} |xi6=∞for allx∈N}.

We claim that Z 6= ∅. Indeed, xN 6= ∞, that is xNi 6= ∞ for some index i.

Supposing that there exists anx∈N withxi=∞givesxNi =∞, a contradiction.

Thereforei∈Z. In fact,

Z={i∈ {1, . . . , n} |xNi 6=∞}. We have

N ={x∈X|xi6=∞for alli∈Z} sincex∈N ⇐⇒ xN∗x=xN ⇐⇒ xi6=∞for alli∈Z.

Comparing our two descriptions ofN, we conclude that for all x∈X, one has λ(x) =∞if and only if there existsi∈Z such that xi =∞. This indicates that ann-ary term function, if it is to extendλ, must involve all of the variables with indices inZ.

LetπZ:Mn →MZ be restriction toZ. It is clear thatπZ(N)⊆MZ. We will now argue thatπZ(N) is a substructure ofM

Z.

Letg ∈G bem-ary and lety1, . . . , ym∈πZ(N). To showgMZ(y1, . . . , ym) is in πZ(N), letx1, . . . , xm∈N be such thatπZ(x1) =y1, . . . , πZ(xm) =ym. Then gX(x1, . . . , xm)∈X and for alli∈Z, we havegX(x1, . . . , xm)i ∈M. Hence, by our previous claim,λ(gX(x1, . . . , xm))6=∞, that isgX(x1, . . . , xm) is inN. Since thexj agree with theyj onZ, we have

gMZ(y1, . . . , ym) =πZ(gX(x1, . . . , xm))∈πZ(N).

Leth∈H be anm-ary partial operation symbol and let y1, . . . , ym ∈πZ(N) with y1, . . . , ym ∈ dom(hMZ). That is, yi1, . . . yim ∈ dom(hM) for each i ∈ Z.

Letx1, . . . , xm ∈N be such thatπZ(x1) =y1, . . . , πZ(xm) =ym. We then have x1∗xN, . . . , xm∗xN ∈ X and again πZ(x1∗xN) = y1, . . . , πZ(xm∗xN) =ym. Checking thatx1∗xN, . . . , xm∗xN ∈dom(hX), as before we conclude

hMZ(y1, . . . , ym) =πZ(hX(x1∗xN, . . . , xm∗xN))∈πZ(N), completing the argument.

Now, supposex, y∈Nagree onZ, that isπZ(x) =πZ(y). Sincex∗xN =y∗xN andλpreserves∗, we have

λ(x) =λ(x)∗λ(xN) =λ(x∗xN) =λ(y∗xN) =λ(y)∗λ(xN) =λ(y).

This shows that the mapλ0Z(N)→M given for allz∈πZ(N) by λ0(z) =λ(x), wherexis any element ofN such that πZ(x) =z,

(8)

is well defined. In fact, one may show that λ0 is a morphism from πZ(N) toM by using an argument similar to the one above and the fact thatλpreserves the∼ (partial) operations and relations inGM∪HM∪RM.

We now have a substructureπZ(N) ofM

Z and a morphism λ0Z(N)→M

∼, so, by invoking (IC) for M, our induced λ0 extends to a Z-ary term function tM:MZ →M. To complete the proof, it remains to show that the n-ary term function

sM = (t∗vi1∗ · · · ∗vik)M, whereZ ={i1, . . . , ik}, extendsλ.

We have already shown that for all x∈ X, one has λ(x) = ∞ if and only if xi=∞for some i∈Z. It is easy to see that for all x∈Mn,

sM(x) =∞ ⇐⇒ xi=∞for somei∈Z, sincesinvolves all variables with indices inZ. Hence

λ(x) =∞ ⇐⇒ sM(x) =∞for allx∈X.

Finally, letx∈N. Sincexi∈M for alli∈Z, we have λ(x) =λ0Z(x)) =tMZ(x)) =tM(x) =tM(x)

=tM(x)∗xi1∗ · · · ∗xik= (t∗vi1∗ · · · ∗vik)M(x) =sM(x),

as required.

The following proposition shows the necessity of the relationM in the structure M∼:=hM;GM∪ {∗} ∪ {∞}, HM, RM∪ {M}, τi,

provided no nullaries appear inGM. We will later provide examples where, in the presence of nullaries, the unary relationM may be avoided.

Proposition 2.4. SupposeGM contains no nullaries. Then GM∪ {∗} ∪ {∞} ∪HM∪RM does not yield a duality on the algebra M∈A=ISPM.

Proof. We will show that the constant map c∞:D(M) → {∞}, although in ED(M), cannot be an evaluation at anya∈M.

Letx∈D(M) =A(M,M) and note that for alla, b∈M, we havex(a) = x(a∗b) = x(a)∗x(b). This shows that if x(b) = ∞ for some b ∈ M, then x(a) =∞for alla∈M, henceD(M) consists precisely of the endomorphisms of M together with the constant mapM→ {∞}.

(9)

Clearly,c∞ ∈ED(M), but any endomorphism ofM evaluated at anya∈M

will differ from∞.

According to Lemma 2.2, we may replaceM with M in Theorem 2.3, pro- vided M has a left-zero term. As an example of when this is not the case, let 2 be the two-element meet-semilattice h{0,1};·i. Then the structure 2

∼ = h{0,1};·,0,1;τidualises2and indeed satisfies (IC) (see [2]). Observe that2can- not have a left-zero term, since the identity u≈ v is satisfied in a semilattice if and only if the terms uandv contain precisely the same variables. Hence, when applying Theorem 2.3, we obtain a duality for the algebra2 =h{0,1,∞};·,∗i. Since∗ is a homomorphism (2)2→2, we have the medial law

(x·y)∗(w·z)≈(x∗w)·(y∗z)

true in2. Using this together with the idempotence of the operations· and ∗, we obtain the distributive laws

x∗(y·z)≈(x∗y)·(x∗z) (x·y)∗z≈(x·z)∗(y·z),

hence2 is a semiring. Such semirings were considered in [16] and [15]

Note that there is a duality for the three-element chain2arising from Propo- sition 2.1 given by the structure 2

=h{0,1,∞}; End (2)∪ {·};τi(see also [6]).

3. P lonka Sums and Regularisations

In this section, we consider a fixedplural typeF. That is,F has an operation symbol of arity at least 2 and no nullaries.

An algebra Mof type F is calledregular if it satisfies only regular identities (that is, identities in which the same variables appear on both sides), andirregu- larotherwise. For a varietyVofF-algebras, theregularisationVofVis defined to be the variety of algebras satisfying all regular identities that are satisfied inV. An alternative description ofVmay be given as follows.

Given a semilattice S = hS;·i, define the F-algebra SF := hS;FSFi having fundamental operations given byfSF(x1, . . . , xn) =x1· · ·xnfor allx1, . . . , xn∈S and each n-ary f ∈ F. Denote by SLF the class consisting of the SF where S ranges over all semilattices. Evidently,SLF is precisely the variety ofF-algebras satisfying all regular identities of type F. Indeed, as before, any SF ∈ SLF

satisfies an identityu≈vof typeF if and only if the term functionsuSF andvSF are the same product of variables inS, that is, if and only if u≈v is a regular identity. We now have

V=HSP(V∪SLF).

(10)

Also, note thatSLF is term equivalent to the variety of semilattices since in any SF we may recover the original·operation inSby: x·y=fSF(x, y, . . . , y) where f ∈F is any operation symbol of arity greater than 1. Accordingly, we callSLF

the variety ofsemilattices of type F.

Recall that a semilatticeS=hS;·imay be regarded as a (small) category (S) with object setS and homsets given by

(S)(s, t) =

{s→t} ifs≤t,

∅ otherwise,

for all s, t∈S (where s≤t ⇐⇒ s·t =s). Note that (S) has products of any two of its objects, the product ofsandt in (S) being s·t.

Definition 3. LetS=hS;·ibe a semilattice and letVbe a variety of algebras of typeF. LetQbe a contravariant functor from (S) toV. We writeAs:=Q(s) for each s ∈ S, and ϕt,s: = Q(s → t) for each s ≤ t in S, the fibre map from At to As. Then the P lonka sum of Qis the F-algebra Awith universe A= ˙S

{As|s∈S} and fundamental operations given, for eachn-ary f ∈F, by fA(x1, . . . , xn) =fAss1,s(x1), . . . , ϕsn,s(xn))

for allx1, . . . , xn ∈A, wherexi∈Asi ands=s1· · ·sn. We will sometimes sayA is the P lonka sum of the system offibres(As|s∈S) with fibre maps (ϕt,s|s≤t) over thesemilattice replica S.

For example, the algebraA in Definition 2.1 is the P lonka sum of the func- tor Qfrom the category corresponding to the two-element meet-semilattice2:=

h{0,1};·i, whereQ(1) =Aand Q(0) is the trivialF-algebra h{∞};Fi. The only non-identity fibre map Q(0→1) :A→ {∞}is constant. Note that1, where1 is the trivialF-algebrah{1};Fi, is isomorphic to2F.

An algebra M is called strongly irregular if it has a left-zero term ∗ (cf.

Lemma 2.2). For example, any algebra with an underlying lattice structure is strongly irregular (we may take the term x∗y to be x∨(x∧y)). Similarly, an algebra with an underlying group structure is strongly irregular (we may take the termx∗y to bexy1y). More generally, any non-trivial algebra in a congruence- modular variety is strongly irregular since amongst the terms of such an algebra is a 4-ary Day termd(which is not a projection) satisfyingd(x, y, y, x)≈x, and we may takex∗y to bed(x, y, y, x).

A variety Vof algebras will be called strongly irregularif there is a binary term∗which is a left-zero term on every algebra inV. It turns out that a strongly irregular varietyVhas a basis for its identities consisting of some regular identities together with the single identityx∗y≈x(see [17] or [11]). The regularisation of Vthen has a very concrete description:

(11)

Theorem 3.1. ([9], [10], [11], [17], [12]) LetVbe a strongly irregular variety of F-algebras defined by a set Σof regular identities and a single identity of the formx∗y≈x. Then the following classes coincide:

(1) The regularisationV ofV;

(2) The class of P lonka sums of algebras inV;

(3) The variety of F-algebras defined by the identities Σ and the following identities (for eachn-aryf ∈F):

x∗x≈x,

(x∗y)∗z≈x∗(y∗z), x∗y∗z≈x∗z∗y,

f(x1, . . . , xn)∗y≈f(x1∗y, . . . , xn∗y), y∗f(x1, . . . , xn)≈y∗x1∗ · · · ∗xn.

Observe that the first three identities of (3) above tell us that for an algebraA inV, the term reducthA;∗iis an idempotent left normal semigroup.

In the presence of strong irregularity, there is also a characterisation of the subdirectly irreducibles inV:

Theorem 3.2. ([8]) Let V be a strongly irregular variety. The subdirectly irreducible members of V are the algebras A and A, where A ranges over all subdirectly irreducible members of V, and the algebra 1, where 1 is a trivial algebra inV.

From the point of view of natural duality theory, this gives a corollary indi- cating that a 2-sorted duality may be necessary if we are seeking to lift a given natural duality for a strongly irregular variety Vto a duality covering the whole regularisationV. We need not haveV=ISPM even whenV=M for someM.

A detailed treatment of multi-sorted dualities may be found in Chapter 7 of [2].

Corollary 3.3. LetVbe a strongly irregular variety withV=ISPMfor some algebra M. Then V=ISP(M,1), and further, V=ISPM if and only if M has a one element subalgebra.

Let A∈ V where V is strongly irregular and satisfies the identity x∗y ≈ x.

Suppose (according to Theorem 3.1) thatAis a P lonka sum of fibres (As|s∈S) over its semilattice replicaSwith fibre maps (ϕs,t|s≥t). We define thecanonical homomorphism µA: A → SF by µA(As) = s. The kernel Φ of µA (whose congruence classes are exactly the fibres ofA) may in fact be recovered via

(x, y)∈Φ ⇐⇒ x∗y=xandy∗x=y.

(12)

Note that the quotient algebra A/Φ is isomorphic to SF. For convenience, we will often identifyµAwith the semigroup homomorphism fromhA;∗itoS=hS;·i having the same definition.

Observe that for s ≥ t in S, for any a ∈ As and b ∈ At we have a∗Ab = ϕs,t(a)∗Atϕt,t(b) =ϕs,t(a)∗Atb=ϕs,t(a). Hence we may recover the fibre map ϕs,tby

ϕs,t(a) =a∗b whereb∈Atis arbitrary for alla∈As.

From the operation∗, we may define a partial order on Aby setting x≤y:⇐⇒ x=y∗x.

Under this ordering, each fibre ofAis an anti-chain. Further, for allx, y, w, z∈A, ifx≤ y thenx∗w≤y∗wand z∗x≤ z∗y, that is,hA;∗i forms a partially ordered semigroup under ≤. Since µA : hA;∗i → S is a homomorphism, it follows immediately that µA is order preserving, that is, x ≤ y in A implies µA(x)≤µA(y) inS.

SupposeA,B∈Vwith semilattice replicasSandTrespectively and letu:A→ B be a map. We may attempt to “define” Γu:S→T, thereplica map ofu, by

Γu(s) =µB(u(x)) wherex∈µA1(s) for alls∈S, although, as it stands, Γu need not be well defined.

The following lemma characterises homomorphisms inV. They are, in a sense,

“P lonka sums” ofV-homomorphisms over a replica map.

Lemma 3.4. Let A,B ∈ V (where V is strongly irregular and satisfies x∗ y ≈x). Suppose A has semilattice replica S, fibres (As|s ∈ S) and fibre maps (ϕs,t|s ≥ t). Suppose B has semilattice replica T, fibres (Bv|v ∈ T) and fibre maps(φv,w|v≥w).

Let u:A→B be a map such that for eachs∈S, the restriction uAs:As→B

is a homomorphism. Then the following are equivalent:

(1) uis a homomorphismA→B;

(2) The replica mapΓofuis a well-defined semilattice homomorphismS→T andupreserves the order≤;

(3) upreserves∗;

(4) Γ is a well-defined semilattice homomorphismS→T and for eachs≥t inS, we have

u(ϕs,t(a)) =φΓ(s),Γ(t)(u(a)) for alla∈As.

(13)

Proof. (1) =⇒ (2): This is a routine check after noting that a homomorphism must preserve the term function∗. (Hence we also have (3) =⇒ (2)).

(2) ⇐⇒ (3): Letu:A→B be an order-preserving map such that the replica map Γ is a well-defined semilattice homomorphism. We obtain

Γ(µA(x)) =µB(u(x))

for all x ∈ A, since certainly x ∈ µA1A(x)). Let x, y ∈ A. Using the above together with the fact thatµA andµBare semigroup homomorphisms, we have

µB(u(x∗y)) =µB(u(x)∗u(y)),

showing thatu(x∗y) andu(x)∗u(y) are in the same fibre of B.

Note thatx∗y≤xhenceu(x∗y)≤ u(x) sinceuis order preserving. Then u(x∗y)∗u(x)∗u(y)≤ u(x)∗u(y), from which, by definition we obtain

u(x∗y)∗[u(x)∗u(y)] = [u(x)∗u(y)]∗u(x∗y).

Hence u(x∗y) =u(x)∗u(y), sinceu(x∗y) andu(x)∗u(y) lie in the same fibre, showing (3). (Note that the proof of (2) ⇐⇒ (3) is independent of our assumption that the restriction ofuto each fibre be a homomorphism.)

(3) =⇒ (4): Again, preservation of ∗ ensures that the replica map Γ is a well-defined semilattice homomorphism. Lets≥tinSanda∈As. Letb∈Atbe arbitrary. Thenu(a)∈BΓ(s)andu(b)∈BΓ(t)with Γ(s)≥Γ(t) inT. Therefore

u(ϕs,t(a)) =u(a∗b) =u(a)∗u(b) =φΓ(s),Γ(t)(u(a)).

(4) =⇒ (1): We must show that for eachn-aryf ∈F and eacha1, . . . , an∈A, we have

u(fA(a1, . . . , an)) =fB(u(a1), . . . , u(an)).

By using the P lonka sum description of the fundamental operations ofA andB given in Definition 3.1 and the assumption that the replica map Γ is a well-defined semilattice homomorphism, this condition becomes

u(fAss1,s(a1), . . . , ϕsn,s(an)) =

fBΓ(s)Γ(s1),Γ(s)(u(a1)), . . . , φΓ(sn),Γ(s)(u(an)))

whereai∈Asi ands=s1· · ·sn inS. This is easily verified since, by assumption, we haveuAs:As→BΓ(s)a homomorphism and u(ϕsi,s(ai)) =φΓ(si),Γ(s)(u(ai))

for eachi.

Starting from a dualisable, strongly irregular algebraM, we may use the general theory of P lonka sums to produce a version of Theorem 2.3 that preserves the type ofMwhile weakening the assumptions on the duality forM. SinceMis strongly irregular, by Lemma 2.2,Mis term equivalent to the algebraMof the previous section. LettingV :=HSPM, we have M ∈ V. Algebras in ISPM (which is then a subclass ofV) are P lonka sums ofV-algebras by Theorem 3.1. The following lemma tightens this description and allows us to bypass (IC).

(14)

Lemma 3.5. Algebras inISPM are P lonka sums of ISPM-algebras.

Proof. Let A∈ISPM. Then, as remarked above, A is a P lonka sum ofV- algebras where V=HSPM. Let As be a non-trivial fibre of Aand let a6=b in As. By the Separation Theorem, there is a homomorphismu: A→Msuch that u(a) 6=u(b). Sinceu preserves ∗ (which is a term function), by Lemma 3.4 the replica map Γu is well defined and thereforeu(As) is a subset of a single fibre of M. But thenu(As)⊆M (since the only other fibre,{∞}, is trivial and we have u(a)6=u(b)). The restriction ofuto As is then a homomorphismAs →M that separatesaandband hence (again by the Separation Theorem) As∈ISPM.

Using Definition 2.3 we obtain a set of algebraic (partial) operations and re- lations GM∪HM ∪RM on M from any set GM ∪HM ∪RM of algebraic (partial) operations and relations onM.

Theorem 3.6. Let M be a strongly irregular finite algebra dualised by the structureM

∼ =hM;GM, HM, RM;τiof finite type. Then

M∼:=hM;GM∪ {∗} ∪ {∞}, HM, RM ∪ {M};τi dualisesM.

Proof. Let A be a finite algebra in ISPM and let λ: D(A) → M

be a morphism. We aim to extend λ to an A-ary term function MA → M, the result will then follow by the First Duality Theorem and the Duality Compactness Theorem.

In the first case, suppose thatλis the constant map onto{∞}. Sinceλpreserves the unary relationM, we must have for eachx∈D(A), ana∈Asuch thatx(a) =

∞. Then, lettingA={a1, . . . , am}, we see that the term function (v1∗· · ·∗vm)M extendsλsince, for allx∈D(A), we have (v1∗· · ·∗vm)M(x) =x(a1)∗· · ·∗x(am) =

∞=λ(x).

Alternatively, suppose the set of homomorphisms N :={x∈D(A)|λ(x)6=∞}

is non-empty. We enumerateN ={x1, . . . , xl} and define xN :=x1∗ · · · ∗xl,

that is, xN(a) =x1(a)∗ · · · ∗xl(a) for alla∈A. Then xN ∈D(A) and, indeed, xN ∈N, for otherwise

λ(xN) =∞ =⇒ λ(x1)∗ · · · ∗λ(xl) =∞ ⇐⇒ λ(xi) =∞for somei, a contradiction.

(15)

Let

Z :={a∈A|x(a)6=∞for allx∈N}.

We claim thatZis non-empty. CertainlyxN 6=∞(for otherwiseλ(xN) =λ(∞) =

∞sinceλpreserves the constant ∞), that is, xN(a)6=∞ for somea∈A. Now, suppose that there exists an x∈N such thatx(a) =∞. ThenxN(a) =x1(a)∗

· · · ∗x(a)∗ · · · ∗xl(a) = ∞, a contradiction. Therefore a ∈ Z and we have in fact shown that Z ⊇ {a∈ A|xN(a) 6=∞}. Conversely, if a∈ Z then xN(a) = x1(a)∗ · · · ∗xl(a) =x1(a) (since all ofx1(a), . . . , xl(a) are inM), hence

Z ={a∈A|xN(a)6=∞}.

Observe that x ∈ N =⇒ xN ∗x = xN. Conversely, let x ∈ D(A) with xN ∗x=xN and suppose that λ(x) =∞. Then λ(xN) = λ(xN ∗x) =λ(xN)∗ λ(x) =λ(xN)∗ ∞=∞, a contradiction. Therefore

x∈N ⇐⇒ xN ∗x=xN.

But, xN ∗x=xN if and only if, for each a∈A, if xN(a) 6=∞then x(a)6=∞. That is,

N ={x∈D(A)|x(a)6=∞for alla∈Z}, from which we obtain

λ(x) =∞ ⇐⇒ there existsa∈Z such thatx(a) =∞.

Suppose now thatAis a P lonka sum with semilattice replicaS, fibres (As|s∈ S) and fibre maps (ϕs,t|s≥t). According to Lemma 3.5, eachAsis in ISPM.

SincexN:A→M is a homomorphism and therefore preserves ∗, the replica map ΓxN:S → 2 is a well-defined semilattice homomorphism by Lemma 3.4.

(Recall that2denotes the two element meet semilattice on {0,1}, the semilattice replica ofM). We then have, for eachs∈S,

ΓxN(s) =

1 ifs≥σ, 0 ifs6≥σ,

for someσ∈S. Hence, for eacha∈Awitha∈As we obtain xN(a)6=∞ ifs≥σ

xN(a) =∞ ifs6≥σ, showing that

Z=[˙

{As|s≥σ}.

(16)

Consider the least fibre Aσ of Z and let πAσ:MA → MAσ be restriction to Aσ. For each x∈ N, the mapπAσ(x) is a homomorphism Aσ →M, that is, πAσ(N) is a subset of the first dual ofAσ with respect toA:=ISPM, in symbols,

πAσ(N)⊆DM(Aσ) :=A(Aσ,M).

We claim that in fact

πAσ(N) =DM(Aσ).

Let y ∈DM(Aσ), that is, lety:Aσ → M be a homomorphism. We must show that y is the restriction to Aσ of some homomorphism A → M in N. Define y:A→M by

y(a) =

y(ϕs,σ(a)) ifa∈As⊆Z,

∞ otherwise.

It may be verified that y satisfies condition (2) of Lemma 3.4 and is therefore a homomorphism A→ M. Noting that ϕσ,σ is the identity Aσ →Aσ, we have πAσ(y) =y. If a∈Z, thena is in some fibre As of A withs ≥σ and we have y(a) = y(ϕs,σ(a)) ∈ M, that is, y(a) 6= ∞ for all a ∈ Z. Therefore y ∈ N, establishing the claim.

We will now show that given a homomorphism inN, its values onZ are com- pletely determined by its values onAσ, and this also determines its image inM

under λ. Let x, y∈ N agree on Aσ, that is πAσ(x) =πAσ(y). We claim that x and y then agree onZ. Let a ∈Z, say a ∈As where s ≥σ. Using the P lonka sum description of the fundamental operations (and therefore term functions) of A, we have

a∗Aϕs,σ(a) =ϕs,σ(a)∗Aσϕs,σ(a) =ϕs,σ(a).

Hence, by definition,ϕs,σ(a)≤ainA. Sincexis order-preserving by Lemma 3.4, we then have x(ϕs,σ(a)) ≤ x(a) in M. But x(a) and x(ϕs,σ(a)) lie in the same fibre ofM, namelyM, and≤ is the antichain order on M, so we obtain x(a) =x(ϕs,σ(a)) and similarlyy(a) =y(ϕs,σ(a)). Hence for alla∈Z we have

x(a) =x(ϕs,σ(a)) =y(ϕs,σ(a)) =y(a), showing thatxandy agree onZ.

Since∗is a left-zero operation onMandxN takes the value∞on the comple- ment ofZ in A, we also have

x∗xN =y∗xN. Therefore

λ(x) =λ(x)∗λ(xN) =λ(x∗xN) =λ(y∗xN) =λ(y)∗λ(xN) =λ(y).

(17)

According to the above, we may unambiguously define the mapλ0Aσ(N)→ M by

λ0(z) =λ(x) for anyx∈N such thatπAσ(x) =z, for allz∈πAσ(N) =DM(Aσ).

We now show thatλ0 is a morphismDM(Aσ)→M

∼ with respect to the original M∼ structure.

Let r ∈ R be n-ary and let z1, . . . , zn ∈ DM(Aσ) = πAσ(N) be such that (z1, . . . , zn) ∈ rDM(Aσ). That is, (z1(a), . . . , zn(a)) ∈ rM for all a ∈ Aσ. Let x1, . . . , xn∈N withπAσ(x1) =z1, . . . , πAσ(xn) =zn. By the above arguments we have πAσ(x1∗xN) =z1, . . . , πAσ(xn∗xN) =zn. Noting that each xi∗xN takes the value∞outside ofZ, we have, for all a∈As⊆A,

(x1∗xN(a), . . . , xn∗xN(a)) = (x1(a)∗xN(a), . . . , xn(a)∗xN(a))

=

(x1s,σ(a)), . . . , xns,σ(a))) ifa∈Z, (∞, . . . ,∞) ifa6∈Z

=

(z1(a), . . . , zn(a)) ifa∈Z, (∞, . . . ,∞) ifa6∈Z.

Then, using the definition of rM, we have (x1∗xN, . . . , xn∗xN)∈rD(A), and sinceλpreserves rM, we have (λ(x1∗xN), . . . , λ(xn∗xN))∈rM. Hence

0(z1), . . . , λ0(zn)) = (λ(x1∗xN), . . . , λ(xn∗xN))∈rM,

but (λ0(z1), . . . , λ0(zn))∈Mn, therefore (λ0(z1), . . . , λ0(zn))∈rM, that isλ0 pre- serves rM. Similar arguments show that λ0 preserves the (partial) operations in GM∪HM, henceλ0 is a morphismDM(Aσ)→M

∼. SinceM

∼ yields a duality onM, by the First Duality Theoremλ0 extends to an Aσ-ary term function MAσ →M, saytM. Consider theA-ary term

s=t∗ Y

aZ

va

where va is the variable term corresponding to the ath projection term function πa andQ

aZ is the|Z|-fold ∗product.

For x∈ D(A) with x6∈ N, we have some a ∈ Z such that x(a) = ∞, hence sM(x) =tM(x)∗ ∞=∞=λ(x). Alternatively, ifx∈N, we have

λ(x) =λ0Aσ(x)) =tMAσ(x)) =tM(x)

=tM(x) =tM(x)∗ Y

aZ

x(a) =sM(x).

Hence theA-ary term function sM extends λ, completing the proof.

(18)

We conclude this section with several examples. Let D=h{0,1};∨,∧ibe the two-element distributive lattice. The extended operations, ∨ and ∧, on the set D:={0,1,∞}are just the join in the three-element chain 0<1<∞and meet in the three-element chain∞<0<1. The variety (= quasi-variety) generated by Dis precisely the variety of distributive bisemilattices (see [7]). Priestley duality for distributive lattices states that D

∼ = h{0,1}; 0,1,≤, τi dualises D. Thus, by Theorem 3.6,D

:=h{0,1,∞};∗,0,1,∞,≤, τidualisesD, where≤is the order on {0,1,∞} whose only non-trivial comparability is 0 < 1. Note that here we may avoid the unary relation D = {0,1} since any map preserving ≤ and the constants 0,1 must preserveD. The dualising structure forDgiven by Gierz and Romanowska in [7] uses a different order, namely the order≤, with∞<0<1, associated with the the meet operation onD. It could be argued that the order

≤which arises from Theorem 3.6 is more natural than ≤ as it is symmetric in its relationship to the operations∨and∧onD.

IfS=hS;·iis a finite semigroup which possesses a left-zero term and is dualised by a structure of finite type, then Theorem 3.6 shows that the semigroupSis also dualised by a structure of finite type. In particular, every finite group, regarded as a semigroup, has a left-zero term. Since finite abelian groups are dualisable (see Davey [6]), it follows that a semigroup obtained by adding a new zero to a finite abelian group is dualisable (see also [13] and [14]). Similarly, semigroups obtained by adding a new zero to certain non-abelian groups, for example dihedral groups of order 2nwithnodd (see Davey and Quackenbush [4]), are dualisable.

Perhaps the simplest strongly irregular variety is the variety LZof left-zero semigroups (algebras with one binary operation∗satisfying the identityx∗y≈x).

This variety is term equivalent to the variety of non-empty sets and we have LZ=ISPLwhereL=h{0,1};∗iis the uniquely determined two-element left zero semigroup on{0,1}. The structure

∼L=h{0,1};∨,∧,0,0,1;τi,

whereh{0,1};∨,∧,0,0,1iis the two element Boolean algebra yields a (strong) du- ality onLZ. Theorem 3.6 gives a duality for the left normal idempotent semigroup L via the structure

L∼=hL;∨,∧,0,∗,0,1,∞;τi

where ∨, ∧ on L are the distributive bisemilattice operations as in the first example and0 is given by:

0 0 1 ∞

1 0 ∞

We may again avoid the unary relation Ldue to the presence of the constants 0,1 and either of the chain orders arising from ∨ or ∧. Further, since ISPL = LZ and L has a one element subalgebra, by Corollary 3.3 we obtain ISPL =

(19)

LZ. By Theorem 3.1, LZ is precisely the variety of left normal idempotent semigroups. This duality is closely related to the regularised Lindenbaum-Tarski duality discussed in [13] and [14].

4. Quasi-regularised Sets

It is natural to ask what may be said about the dualisability of other kinds of P lonka sums with dualisable fibres. Here, we consider a two-fibred P lonka sum where we attach the left-zero bandLto the bottom of the semilattice and attach a trivial fibre to the top and give a “bare-hands” proof that the resulting semigroup is dualisable.

ConsiderL0 given by the following table:

∗ 0 1 10

0 0 0 0

1 1 1 1

10 1 1 10

L0 ∈LZis the P lonka sum of the functorQfrom the categorical two-element meet semilattice (2), where Q(0) = L while Q(1) is the trivial idempotent semigroup h{10};∗i and the fibre mapφ1,0 =Q(0→1) :{10} →L distinguishes the element 1. (Ifφ1,0 instead distinguishes 0, the semigroup obtained is isomorphic toL0.)

The dualising structure forL0 will contain a modification of L

∼consistent with the P lonka sum construction, just as we obtained L

. However, in contrast with

∼L, we will need to add an essentially “new” binary relation. The modified operations ∨0, ∧0 and 00 on L0 obtained from ∨, ∧ and 0 on L are given by the following tables:

0 0 1 10

0 0 1 1

1 1 1 1

10 1 1 10

0 0 1 10

0 0 0 0

1 0 1 1

10 0 1 10

00 0 1 10

1 0 0

A routine check ensures that these operations are algebraic, as is the binary relation

%:={(0,1),(1,1),(1,10)},

which, as a subalgebra of (L0)2, is isomorphic toL0. We also include the semigroup operation∗and the unary relation L, hence

L0

∼:=hL0;∗,∨0,∧0,00,0,1,10;%, L;τi will be our dualising structure.

(20)

Theorem 4.1. L0

∼ dualisesL0. Proof. SinceL0

∼is of finite type, it suffices to show that (IC) holds by the Duality Compactness Theorem. To this end, letn∈N, let X be a (closed) substructure of (L0

∼)n, and letα: X→L0

∼ be a morphism.

LetT :={1,10}and note thathT;∗i is a meet-semilattice with 1<10.

Observe thatLn, with the operations∨0,∧0,00suitably restricted together with the constants 0,1, forms a Boolean algebra havingX∩Lnas a subalgebra. Sinceα preserves the unary relationL, the restriction ofαtoX∩Ln is a Boolean algebra homomorphism onto the two-element Boolean algebra on{0,1}. Therefore

αXLn(x) =

1 ifx≥β, 0 ifx6≥β,

for all x∈X ∩Ln and some atom β of X∩Ln. Here, ≤is the usual order on {0,1} extended pointwise.

We may similarly characteriseαrestricted to the (non-empty) semilatticehX∩ Tn;∗i ≤ hTn;∗i. Since it is a semilattice homomorphism onto the two element meet semilatticehT;∗i, we have

αXTn(x) =

10 ifx≥σ, 1 ifx6≥σ,

for all x ∈ X ∩Tn and some σ ∈ X ∩Tn with σ 6= 1 (since α preserves the nullary 1).

Our first claim is that the set of indices

{i∈ {1, . . . , n} |βi= 1 andσi= 10}

is non-empty. Suppose to the contrary that for alli∈ {1, . . . , n}, wheneverβi= 1, we haveσi= 1. Thenσi = 1 wheneverβ00i= 0, from which we obtainβ00%σ. But thenα(β)00%α(σ) sinceαpreserves%and00, which gives 0%10, a contradiction.

Fix aj∈ {i∈ {1, . . . , n} |βi= 1 andσi = 10} and let K:={i∈ {1, . . . , n} |σi= 10 andi6=j}. Now, lettbe the term function (L0)n→L0 given by

t(x) :=xj∗(Y

iK

xi)

for allx∈(L0)n. We will show thattextends α.

(21)

Clearly,tsatisfies the following for allx∈(L0)n: t(x) = 0 ⇐⇒ xj= 0;

t(x) = 1 ⇐⇒ xj6= 0 andxi6= 10 for some i∈K∪ {j}; t(x) = 10 ⇐⇒ xi= 10 for alli∈K∪ {j}.

Letx∈X. We observe that 10∗x∈X∩Tn and hence α(x) = 10 ⇐⇒ 10∗α(x) = 10

⇐⇒ α(10∗x) = 10

⇐⇒ 10∗x≥σ

⇐⇒ (10∗x)i= 10 for alli∈K∪ {j}

⇐⇒ xi= 10 for alli∈K∪ {j}

⇐⇒ t(x) = 10.

Sincex∗0∈X∩Ln, we have

α(x) = 0 ⇐⇒ α(x)∗0 = 0

⇐⇒ α(x∗0) = 0

⇐⇒ x∗06≥β

⇐⇒ (x∗0)j= 0 (see note below)

⇐⇒ xj= 0

⇐⇒ t(x) = 0.

To see thatx∗06≥β implies (x∗0)j= 0 in the above argument, suppose that x∗06≥β but (x∗0)j= 1. Then (x∗0)i= 0 andβi= 1 for somei6=j. But then, sinceβj= 1, inX∩Ln we have 0<(x∗0)∧β < β, contradicting the fact thatβ

is an atom.

We close by showing that ISPL0 is in fact the quasi-regularisation of the variety LZ. The quasi-regularisation of a variety V of type F, as introduced by Bergman and Romanowska in [1], is defined to be the quasi-variety generated by V∪SLF, in symbols: Q(V∪SLF). If V is strongly irregular, it is shown in [1]

that the quasi-regularisation ofVis always a proper subclass of its regularisation V. The following Theorem summarises the characterisation ofQ(V∪SLF) given there.

Theorem 4.2. ([1]) Let A be an algebra in the regularisation of a strongly irregular varietyV(in which the identityx∗y≈xis satisfied). Assume thatAis the P lonka sum of subalgebras (As|s∈S)over the semilattice S, with fibre maps ϕs,t. The following are equivalent.

(22)

(1)A∈ISP(V∪SLF);

(2)For every s≥t in S, the homomorphismϕs,tis injective;

(3)A satisfies any of the three equivalent quasi-identities given below:

(a)

x∗y≈x y∗x≈y x∗z≈z∗x≈z y∗z≈z∗y≈z

=⇒ x≈y;

(b)

x∗y≈x y∗x≈y x∗z≈z y∗z≈z

=⇒ x≈y;

(c) x∗z≈y∗z =⇒ x∗y≈y∗x.

(4)A∈Q(V∪SLF).

Proof. We show here that the quasi-identities in (3) are equivalent, and refer the reader to [1] for a proof of the equivalence of (1) to (4), where the quasi-identity (a) is denotedq.

LetA∈Vhave semilattice replicaS, canonical homomorphismµ:A→Sand fibre maps (ϕt,s|s≤t).

Let Asatisfy (a). To show that Asatisfies (b), it will suffice to show that if x, z ∈A and x∗z=z, then z∗x=z. Ifx∗z =z, then by definition we have z≤xand consequentlyµ(z)≤µ(x) inS. But thenz∗x=z∗ϕµ(x),µ(z)(x) =z.

Now, let Asatisfy (b) and let x, y, z∈Awith x∗z=y∗z =w, say. Clearly x∗w=y∗w=w. We have (x∗y)∗(y∗x) =x∗y and (y∗x)∗(x∗y) =y∗x.

Also, (x∗y)∗w=x∗w=wand similarly (y∗x)∗w=w, hencex∗y=y∗xby (b), showing thatAsatisfies (c).

Let A satisfy (c) and letx, y, z ∈ A satisfy the antecedent of (a). Using (c), fromx∗z=y∗z=z we obtainx∗y =y∗x, butx∗y=xandy∗x=y shows thatxandy are in the same fibre, from which it follows thatx=y.

We will denote the quasi-regularisation ofLZ byLZq. To show that LZq = ISPL0, we will rely on Lemma 3.4, characterising the homomorphisms in LZ. Interestingly, a LZq-semigroup will be seen to be not only a disjoint union of left-zero semigroups, but simultaneously a disjoint union of semilattices!

Theorem 4.3. LZq =ISPL0.

Proof. We need only showLZq ⊆ISPL0, the reverse inclusion being given by Theorem 4.2. Let A∈ LZq have semilattice replicaS and canonical homomor- phism µ:A → S. Let a, b ∈ A with a 6= b. We must find a homomorphism A→L0 that separates aandb. Firstly, observe that the semilattice replica ofL0 is the two element meet semilattice on{0,1}, and we have 1<10.

参照

関連したドキュメント

As with subword order, the M¨obius function for compositions is given by a signed sum over normal embeddings, although here the sign of a normal embedding depends on the

Keywords and Phrases: moduli of vector bundles on curves, modular compactification, general linear

Likewise we show that any decomposition of the complete graph into strongly regular graphs of (negative) Latin square type is an amorphic association scheme.. We study strongly

It is suggested by our method that most of the quadratic algebras for all St¨ ackel equivalence classes of 3D second order quantum superintegrable systems on conformally flat

Following Speyer, we give a non-recursive formula for the bounded octahedron recurrence using perfect matchings.. Namely, we prove that the solution of the recur- rence at some

For arbitrary 1 &lt; p &lt; ∞ , but again in the starlike case, we obtain a global convergence proof for a particular analytical trial free boundary method for the

Next, we will examine the notion of generalization of Ramsey type theorems in the sense of a given zero sum theorem in view of the new

We show that a non-symmetric Hadamard spin model belongs to a certain triply regular Bose-Mesner algebra of dimension 5 with duality, and we use this to give an explicit formula for