Applied Categorical Structures 6: 87–103, 1998.
c 1998 Kluwer Academic Publishers. Printed in the Netherlands.
87
Categorical Aspects in Projective Geometry
CLAUDE-ALAIN FAURE and ALFRED FRÖLICHER
Section de Mathématiques, Université de Genève, Case postale 240, CH-1211 Genève 24,
Switzerland
(Received: 8 December 1995; accepted: 5 June 1996)
Abstract. After introducing morphisms between projective geometries, some categorical questions
are examined. It is shown that there are three kinds of embeddings and two kinds of quotients.
Furthermore the morphisms decompose in a canonical way into four factors.
Mathematics Subject Classifications (1991). 18B99, 51A10.
Key words: morphisms and homomorphisms of projective geometries, subgeometries and subspaces, quotient geometries, initial and final morphisms, sections and retractions, decomposition
of a morphism.
Introduction
Categorical questions in projective geometry are new, since morphisms between
projective geometries have been introduced only recently; cf. [3, 4] and [5]. Many
questions about the category P roj of projective geometries ought to be studied;
in this paper we examine three of these topics, namely embeddings, quotients
and factorizations of morphisms.
Since a morphism between projective geometries is a map between the underlying point sets which in general is not defined for all points of the source, the
natural forgetful functor has its values in the so-called partial-map category P ar.
The objects of P ar are the sets, and the morphisms from X to Y are the partial
maps, i.e. the maps f : X \ N → Y where N ⊆ X . We call N the kernel of f
and we write f : X− → Y and N = Ker f . P ar is known to be equivalent to
the category Set⋆ of pointed sets; cf. [1].
By an embedding we understand a monomorphism which is initial with respect
to the forgetful functor P roj → P ar. The embeddings turn out to be, up to
isomorphy, the inclusions of the usual (projective) subgeometries. But subgeometries can have undesired features. For instance the dimension of a subgeometry
can be bigger than the one of the ambiant geometry. It will be shown as an
example that the usual real projective plane contains a subgeometry of dimension d for any d 6 ℵ0 . Several equivalent geometric conditions which prevent
such a pathological behavior will be given. The respective embeddings will be
called proper embeddings. Still smaller is the class of the embeddings which
VTEX(ZJ) PIPS No.: 115592 MATHKAP
APCS232.tex; 5/02/1998; 13:54; v.7; p.1
88
C.-A. FAURE AND A. FRÖLICHER
are isomorphic to the inclusion of a subspace. These turn out to be exactly the
sections of the category P roj .
By a quotient we understand an epimorphism which is final with respect to
the functor P roj → P ar. Among these one finds as special case the quotients
by a given subspace which were considered in [4]; they are retractions of P roj .
Every projective geometry is shown to be a quotient of the coproduct of its lines.
An example will show that three points of a quotient geometry can be collinear
even if no collinear preimages exist.
In many categories the morphisms can be factorized in two special morphisms.
For the category of projective geometries one even gets a factorization into four
morphisms.
1. Morphisms of Projective Geometries
In [4] we defined a projective geometry as a set G (called the point set) together
with a ternary relation ℓ on G (called collinearity) satisfying the following three
axioms:
(L1 ) ℓ(a, b, a) ∀a, b ∈ G;
(L2 ) ℓ(a, p, q), ℓ(b, p, q) and p 6= q ⇒ ℓ(a, b, p);
(L3 ) ℓ(p, a, b) and ℓ(p, c, d) ⇒ ∃q with ℓ(q, a, c) and ℓ(q, b, d).
A subspace of G is a subset E ⊆ G with the property a 6= b ∈ E and ℓ(a, b, c) ⇒
c ∈ E . A subgeometry of G is a subset G′ ⊆ G such that G′ together with the
restriction of ℓ to G′ is itself a projective geometry.
NOTATIONS 1.1. Let G be a projective geometry, A ⊆ G and a, b ∈ G;
1◦ C(A) will denote the smallest subspace of G containing A;
2◦ S(A) will denote the smallest subgeometry of G containing A;
3◦ a ⋆ b := C{a, b}.
The subspaces of the form a ⋆ b where a 6= b are called lines; obviously a ⋆ a =
{a}.
There exist many equivalent axiomatic definitions of projective geometries, in
particular by means of points, lines and an incidence relation (cf. e.g. [7]). We
emphasize that we do not require that G is irreducible, i.e. that every line of
G contains at least three points. G is called arguesian if it is irreducible, of
dimension > 2 and has Desargues’ property.
DEFINITION 1.2. A morphism from a projective geometry G1 to a projective
geometry G2 is a map g: G1 \ N → G2 defined on the complement of a subset
N ⊆ G1 and satisfying the following axioms:
(M1 ) N is a subspace of G;
APCS232.tex; 5/02/1998; 13:54; v.7; p.2
CATEGORICAL ASPECTS IN PROJECTIVE GEOMETRY
89
(M2 ) If a, b ∈
/ N, c ∈ N and a ∈ b ⋆ c, then ga = gb;
(M3 ) If a, b, c ∈
/ N and a ∈ b ⋆ c, then ga ∈ gb ⋆ gc.
The set N is called the kernel of g and will be denoted by Ker g; the set
g(G1 \ N ) is called the image of g and will be denoted by Im g. The morphism
g is called constant if | Im g| 6 1 and empty if Im g = ∅ (i.e. if Ker g = G1 ). If
g is a morphism from G1 to G2 we write g: G1 − → G2 , or also g: G1 → G2
if one knows that Ker g = ∅.
A morphism g: G1 − → G2 is called a homomorphism if it satisfies also the
axioms
(M4 ) If a, b ∈
/ N and ga 6= gb, then ga ⋆ gb ⊆ g(a ⋆ b);
(M5 ) If a, b ∈
/ N, a 6= b and ga = gb, then (a ⋆ b) ∩ N 6= ∅.
One easily verifies that the composite of two morphisms in the sense of the above
definition is a morphism. Hence one obtains a category; it will be noted P roj .
Similarly, the projective geometries together with the homomorphisms form a
subcategory of P roj .
The axioms (M1 ), (M2 ), (M3 ) together are equivalent with each of the following axioms:
(M) If F is a subspace of G2 , then g−1 F ∪ N is a subspace of G1 ; cf. 3.1.1
in [4];
f
(M) g(C1 A \ N ) ⊆ C2 (g(A \ N )) for each A ⊆ G1 .
Furthermore one has for any morphism g: if a, b ∈
/ Ker g and ga 6= gb, then
g(a ⋆ b) ⊆ ga ⋆ gb and hence g(a ⋆ b) = ga ⋆ gb if (M4 ) holds.
PROPOSITION 1.3. A partial map g: G1 \ N → G2 between projective geometries Gi is a morphism if and only its restriction to every line δ of G1 is a
morphism δ − → G2 .
Proof. The restriction of g to δ is the composite g ◦ jδ , where jδ : δ → G1
is the inclusion. The stated condition is trivially necessary. Conversely, suppose
that for every line δ ⊆ G1 the map g ◦ jδ : δ \ (N ∩ δ) → G2 is a morphism. Let
F ⊆ G2 be a subspace. One has (g−1 F ∪ N ) ∩ δ = (g−1 F ∩ δ) ∪ (N ∩ δ) =
jδ−1 (g−1 F ) ∪ (N ∩ δ) = (g ◦ jδ )−1 F ∪ (N ∩ δ). On the right-hand side one has
by (M) a subspace of G1 . Hence the intersection of g−1 F ∪ N with every line
δ is a subspace of G1 and this implies that g−1 F ∪ N is a subspace of G1 . So g
2
satisfies (M) and therefore is a morphism.
PROPOSITION 1.4. The category P roj has arbitrary coproducts.
Proof. Let Gi , i ∈ I be a family of projective geometries. We
may assume that
S
the point sets Gi are pairwise disjoint and consider on G := i∈I Gi the relation
ℓ defined as follows: ℓ(x, y, z) holds if there exists i ∈ I such that x, y, z ∈ Gi
and ℓi (x, y, z) or if x, y, z are not all three different. One easily verifies that
APCS232.tex; 5/02/1998; 13:54; v.7; p.3
90
C.-A. FAURE AND A. FRÖLICHER
G together with ℓ is a projective geometry, that the inclusions ik : Gk → G are
morphisms, and that the coproduct-property holds. Obviously G is not irreducible
if at least two of the geometries are non-empty.
2
The two kinds of morphisms between projective geometries correspond to two
kinds of morphisms between vector spaces.
DEFINITION 1.5. Let V, W be vector spaces over (skew-) fields K, L. We call
a map f : V → W semilinear (respectively quasilinear) if it is additive and if
there exists a homomorphism (respectively isomorphism) of fields σ : K → L
such that one has f (αx) = σ(α) · f (x) for all α ∈ K, x ∈ V . The category
formed by the vector spaces as objects and the semilinear maps as morphisms
will be denoted by V ec.
To every vector space V one associates a projective geometry PV as follows:
The points of PV are the 1-dimensional subspaces of V ; and three of them
are collinear if they lie in a subspace of dimension 6 2. Every point of PV is
determined (“represented”) by any of its non-zero vectors. Every semilinear map
f : V →W induces a morphism Pf : PV − → PW whose kernel Ker(Pf ) is
formed by the 1-dimensional subspaces of V which lie in ker f . If f is quasilinear,
then Pf is a homomorphism. For the so obtained functor P: V ec → P roj the
following result, called generalized fundamental theorem of projective geometry,
holds. For a proof we refer to [5] and [6].
THEOREM 1.6. Let g: G1 − → G2 be a morphism of projective geometries.
1◦ If G1 , G2 are arguesian there exist isomorphisms ui : Gi → PVi . If in addition not all points of Im g are collinear, then there exists for given u1 , u2 a
semilinear map f : V1 → V2 such that g = u−1
2 ◦ Pf ◦u1 . We will then shortly
∼
write g = Pf .
2◦ If g is non-constant, then the equation g = u−1
2 ◦ Pf ◦ u1 determines f up to
a factor κ ∈ K2 , and f is quasilinear if and only if g is a homomorphism.
PROPOSITION 1.7. For a morphism m: G1 − → G2 of projective geometries
the following conditions are equivalent:
1◦ m is a monomorphism of P roj ;
2◦ Ker m = ∅ and the map m: G1 → G2 is injective.
Proof. 1 ⇒ 2. Suppose that there would exist a point p ∈ Ker m. One chooses any non-empty projective geometry G and considers as g1 : G → G1 the
morphism with constant value p (and empty kernel) and as g2 : G − → G1 the
empty morphism. Then m ◦ g1 = m ◦ g2 but g1 6= g2 , in contradiction with the
hypothesis. Hence Ker m = ∅.
Suppose now that there would exist points a 6= b with ma = mb. One
chooses any non-empty projective geometry G and considers as g1 : G → G1
APCS232.tex; 5/02/1998; 13:54; v.7; p.4
CATEGORICAL ASPECTS IN PROJECTIVE GEOMETRY
91
the morphism having the constant value a and as g2 : G → G1 the morphism
with constant value b. One obtains the same contradiction as before and hence
injectivity of m is proved.
2 ⇒ 1. Let m ◦ g1 = m ◦ g2 . Then, using the hypothesis Ker m = ∅, one
gets Ker g1 = Ker(m ◦ g1 ) = Ker(m ◦ g2 ) = Ker g2 =: N . Furthermore, for
x ∈ G \ N one obtains from mg1 x = mg2 x by injectivity of m the equality
2
g1 x = g2 x. Hence g1 = g2 .
PROPOSITION 1.8. For a morphism e: G1 − → G2 of projective geometries
the following conditions are equivalent:
1◦ e is an epimorphism of P roj ;
2◦ e is surjective, i.e. Im e = G2 .
Proof. 1 ⇒ 2. Suppose that a point p ∈ G2 \ Im e would exist. One chooses
a projective geometry G containing exactly one point q and considers the map
g1 : G2 → G (with Ker g1 = ∅) and the map g2 : G2 \ {p} → G. They both are
(constant) morphisms. In order to show that g1 ◦ e = g2 ◦ e it is enough to verify
that the kernels are the same: Ker(gi ◦ e) = Ker e ∪ e−1 (Ker gi ) = Ker e for
i = 1, 2. Since g1 6= g2 one has a contradiction with the hypothesis and hence e
is surjective.
2 ⇒ 1. Let g1 , g2 be morphisms with g1 ◦ e = g2 ◦ e. Suppose x ∈ Ker g1 .
Choose y ∈ G1 \ Ker e with x = ey . Then y ∈ Ker(g1 ◦e) = Ker(g2 ◦e) = Ker e∪
e−1 (Ker g2 ). Therefore y ∈ e−1 (Ker g2 ) and hence x ∈ Ker g2 . By symmetry one
obtains Ker g1 = Ker g2 . Finally, if x ∈
/ Ker g1 = Ker g2 one chooses z such that
x = ez . Then g1 x = (g1 ◦ e)z = (g2 ◦ e)z = g2 x and g1 = g2 follows.
2
2. Embeddings
PROPOSITION 2.1. For a morphism g: G1 − → G2 of projective geometries
the following conditions are equivalent:
1◦ g is an embedding, i.e. an initial monomorphism of the category P roj ;
2◦ g is a monomorphism and ℓ2 (gx, gy, gz) ⇔ ℓ1 (x, y, z);
3◦ Ker g = ∅ and if a1 , a2 resp. a1 , a2 , a3 are independent points of G1 , then
ga1 , ga2 resp. ga1 , ga2 , ga3 are independent points of G2 ;
4◦ g is an isomorphism of G1 onto a subgeometry G′ of G2 , i.e. one has g = j ◦u
where j : G′ → G2 is the inclusion of a subgeometry and u: G1 → G′ an
isomorphism;
5◦ g is an initial morphism of the category P roj .
Proof. 1 ⇒ 2. Suppose that one has points x, y, z ∈ G1 with ℓ2 (gx, gy, gz).
One chooses as G a projective geometry consisting of three different collinear
points a, b, c and considers the map h: G → G1 sending them respectively
APCS232.tex; 5/02/1998; 13:54; v.7; p.5
92
C.-A. FAURE AND A. FRÖLICHER
to x, y, z . Then g ◦ h: G → G2 is a morphism since the images of a, b, c are
collinear. This implies by hypothesis that h: G → G1 is a morphism and therefore
ℓ(a, b, c) ⇒ ℓ1 (x, y, z). The converse is trivial.
2 ⇒ 3. This follows from the fact that in a projective geometry two points
are independent iff they are different and three points are independent iff they
are non-collinear.
3 ⇒ 4. The map g: G1 → G2 can be decomposed as g = j ◦ u where
u: G1 → g(G1 ) is the map ux := gx and j : g(G1 ) → G2 the inclusion map.
Since g is injective, u is bijective. And since one gets ℓ1 (x, y, z) ⇔ ℓ2 (ux, uy, uz)
it follows that G′ := g(G1 ) is a subgeometry of G2 and u an isomorphism.
4 ⇒ 5. Any inclusion j of a subgeometry is trivially an initial morphism.
Therefore also j ◦ u since u is an isomorphism.
5 ⇒ 1. (a) Suppose that there would exist a point p ∈ Ker g. One chooses a
projective geometry G with a subset A which is not a subspace and considers
the map h: G \ A → G1 having the constant value p. Then g ◦ h is the empty
morphism, but h is not a morphism since its kernel is not a subspace. This
contradicts the hypothesis and hence Ker g = ∅.
(b) Suppose now that there would exist points a 6= b with ga = gb. One
chooses G, A as before and considers the map h: G → G1 defined by hx := a
if x ∈ A and hx := b if x ∈
/ A. Then the map g ◦ h: G → G2 is constant and
thus a morphism, but h is not a morphism since h−1 a = A is not a subspace of
G. This contradicts the hypothesis and hence g is injective. By 1.7 one concludes
that g is a monomorphism.
2
PROPOSITION 2.2. For a monomorphism g: G1 → G2 of P roj the following
conditions are equivalent:
1◦
2◦
3◦
4◦
6◦
There exists a basis B of G1 such that gB ⊆ G2 is independent;
A ⊆ G1 independent ⇒ gA ⊆ G2 independent;
For the rank r(E) of any subspace E ⊆ G1 one has: r(E) = r(C2 (gE));
For every subspace E ⊆ G1 one has: E = g−1 C2 (gE);
For every subspace E ⊆ G1 there exists a subspace E2 ⊆ G2 with E =
g−1 E2 ;
5◦ For every subset A ⊆ G1 one has: C1 (A) = g−1 C2 (gA).
Proof. 1 ⇒ 2. We may assume that A is finite. Then there exists a finite
subset B ′ of B with A ⊆ C1 B ′ . By the Steinitz exchange theorem there exists a
basis of C1 B ′ of the form A ∪˙ B ′′ with B ′′ ⊆ B ′ . Then gB ′ ⊆ g(C1 (A ∪˙ B ′′ )) ⊆
C2 (gA ∪˙ gB ′′ ). Since the set gB ′ is independent and since |gA ∪˙ gB ′′ | = |A ∪˙ B ′′ |
= |B ′ | = |gB ′ | < ∞, the preceding inclusion shows that gA ∪˙ gB ′′ and hence
also gA is independent.
e ) one gets C2 (gA) ⊆ C2 (gC1 A) ⊆
2 ⇒ 3. Let A be a basis of E . By (M
C2 C2 (gA) and hence C2 (gE) = C2 (gA). Being independent, gA is therefore a
basis of C2 (gE) and so one gets r(C2 (gE)) = |gA| = |A| = r(E).
APCS232.tex; 5/02/1998; 13:54; v.7; p.6
CATEGORICAL ASPECTS IN PROJECTIVE GEOMETRY
93
3 ⇒ 4. Let E be a subspace of G1 . Then F := g−1 C2 (gE) is also a subspace
of G1 . Since E ⊆ g−1 gE ⊆ g−1 C2 (gE) = F and gF ⊆ C2 (gE) one has
C2 (gE) = C2 (gF ) and hence r(C2 (gE)) = r(C2 (gF )). By hypothesis 3 this is
equivalent with r(E) = r(F ). From this and from E ⊆ F the assertion E = F
follows immediately in the case where r(E) < ∞. The case r(E) = ∞ then
follows also, since the operator C2 is finitary.
4 ⇒ 5. E2 := C2 (gE) is a subspace and by the hypothesis one gets E =
−1
g E2 .
5 ⇒ 6. From A ⊆ g−1 gA ⊆ g−1 C2 (gA) one deduces C1 A ⊆ g−1 C2 (gA).
Furthermore one has by hypothesis C1 A = g−1 E2 for some subspace E2 of
G2 . From A ⊆ g−1 E2 follows gA ⊆ E2 and hence C2 (gA) ⊆ E2 and finally
g−1 C2 (gA) ⊆ g−1 E2 = C1 A.
6 ⇒ 1. Let B be a basis of G1 . Proceeding indirectly we suppose that gB
is dependent, i.e. that there exists b ∈ B such that gb ∈ C2 (gB \ gb). Then
g−1 C2 g(B \ b) = C1 (B \ b), and this contradicts the independence of B .
2
DEFINITION 2.3. 1◦ A proper embedding is a monomorphism which satisfies
the conditions of the preceding proposition (remark that every proper embedding
is an embedding; and that an embedding is proper iff it is an embedding for the
category of closure spaces).
2◦ A subspace-embedding is a proper embedding for which the image is a subspace.
PROPOSITION 2.4. For a morphism g: G1 − → G2 of P roj the following
conditions are equivalent:
1◦ g is a homomorphism with empty kernel;
2◦ g is a subspace-embedding;
3◦ g is a section of the category P roj .
Proof. 1 ⇒ 2. From Ker g = ∅ and (M5 ) one deduces that g is injective and
hence a monomorphism. Hence one gets E = g−1 gE for any subspace E of G1 .
Since (M4 ) implies that gE is a subspace of G2 , condition 5 of 2.2 holds and
hence g is a subspace-embedding.
2 ⇒ 3. By 4 of 2.1 one has g = j ◦ u where j : g(G1 ) → G2 is the inclusion
map and u is an isomorphism. Moreover, g(G1 ) is a subspace of G2 and it is
known, cf. 6.1.3 in [4], that inclusions of subspaces are sections of P roj . With
j also j ◦ u is a section.
3 ⇒ 1. The equation r ◦ g = IdG1 implies Ker g = ∅ and hence also (M5 )
holds. In order to verify (M4 ), let ga 6= gb and z ∈ ga⋆gb. Then rz ∈ r(ga⋆gb) ⊆
rga ⋆ rgb = a ⋆ b and hence z ′ := grz ∈ g(a ⋆ b). From ga 6= gb and rz = rz ′
one deduces by (M3 ) that z = z ′ and hence z ∈ g(a ⋆ b).
2
APCS232.tex; 5/02/1998; 13:54; v.7; p.7
94
C.-A. FAURE AND A. FRÖLICHER
COROLLARY 2.5. Let f : V → W be a semilinear map and g := Pf the
associated morphism of projective geometries.
1◦ g is an embedding of P roj if and only if for 1 6 n 6 3 one has x1 , . . . , xn
linearly independent in V ⇒ f x1 , . . . , f xn linearly independent in W;
2◦ g is a proper embedding of P roj if and only if for all n ∈ N one has
x1 , . . . , xn linearly independent in V ⇒ f x1 , . . . , f xn linearly independent
in W .
Proof. One uses that points of PV are independent if and only if representing
vectors are linearly independent.
2
If f is injective and quasilinear the above conditions hold. We give a more precise
result:
PROPOSITION 2.6. Let, as before, f : V → W be a semilinear map and g :=
Pf . If dim(V ) > 2, then the following conditions are equivalent:
1◦ g is a subspace-embedding (i.e. a section of P roj , cf. 2.4);
2◦ f is injective and quasilinear;
3◦ f is section of V ec.
Proof. 1 ⇒ 2. Since PV contains at least two points, the morphism g is
non-constant. So one can use 2 of 1.6 in order to obtain the quasilinearity of f .
2 ⇒ 3. One verifies that f (V ) is a vector-subspace of W and that f : V →
f (V ) is an isomorphism of V ec. The claim follows since the inclusion of a
vector-subspace obviously is a section of V ec.
3 ⇒ 1. This follows trivially from 2.4
2
EXAMPLE 2.7. An embedding of P(Q(N) ) into P(R3 ).
One chooses α, β, γ ∈ R such that for every N ∈ N one has
aijk ∈ Q and
N
X
aijk αi β j γ k = 0 ⇒ aijk = 0 for all i, j, k
i,j,k=1
and one considers the map f : Q(N) → R3 defined as follows:
f (x1 , x2 , x3 , . . .) :=
X
i
i
xi α ,
X
j
j
xj β ,
X
xk γ
k
k
.
This map f is obviously semilinear. We show that it satisfies condition 1◦ of 2.5.
For n = 1 the condition means injectivity of f which certainly holds.
Suppose now that x, y are vectors of Q(N) such that f x, f y are linearly dependent. Then
P
P
X xi xj
xi αi j xj β j
i
P
αi β j = 0.
j =
i P
i yi α
j
yj β
i,j
yi yj
APCS232.tex; 5/02/1998; 13:54; v.7; p.8
CATEGORICAL ASPECTS IN PROJECTIVE GEOMETRY
95
This implies that
xi xj
= 0 for all
yi yj
i, j
and this is exactly the condition for x, y to be linearly dependent in Q(N) . Let
finally x, y, z be three vectors in Q(N) such that f x, f y, f z are linearly dependent.
Then one has
P
P
P
x γk
x βj
x αi
X xi xj xk
Pi i i Pj j j Pk k k
yi yj yk αi β j γ k = 0.
y
γ
y
β
y
α
=
Pi i i Pj j j Pk k k
i zi α
j zj β
k zk γ
i,j,k
zi zj zk
Since this implies that all the determinants in the sum are zero, the vectors
x, y, z ∈ Q(N) are linearly dependent. By 1 of 2.5 one now deduces that Pf :
P(Q(N) ) → P(R3 ) is an embedding of P roj . The image is a subgeometry of
P(R3 ) which is isomorphic to P(Q(N) ). Obviously the embedding cannot be
proper.
As an immediate consequence one gets: For every n ∈ N there exists a subgeometry of P(R3 ) which is isomorphic to P(Qn ). For n > 3 such an embedding
cannot be proper.
EXAMPLE 2.8. P(Qn ) is isomorphic to a proper subgeometry of P(Rn ).
One chooses as f : Qn → Rn the natural inclusion and uses 2 of 2.5. Obviously
the embedding is not a subspace-embedding. More general examples of proper
embeddings are obtained by choosing any extension of fields K ⊆ L and as f
the natural inclusion K (N) → L(N) or K N → LN .
PROPOSITION 2.9. Let W be a vector space over L and G ⊆ P(W ) an irreducible subgeometry of dimension > 2. Then there exists a subfield K of L and
a vector subspace V of WK , where WK is W considered as vector space over
K , such that G ∼
= P(V ).
Proof. G is arguesian since it is irreducible and inherits from P(W ) the
property of Desargues. Let j : G → P(W ) be the inclusion. Since Im j contains
non-collinear points, there exist by the Fundamental Theorem 1.6 a vector space
V ′ over some field K ′ and a semilinear map f ′ : V ′ → W such that G ∼
= P(V ′ )
′
′
∼
and j = Pf . Using that f is injective one easily shows that V := f ′ (V ′ )
is a vector subspace of WK for K := σ(K ′ ) and that f ′ can be factorized as
f ′ = i ◦ f where i: V → W is the inclusion and f : V ′ → V an isomorphism of
2
V ec. So one has G ∼
= P(V ′ ) ∼
= P(V ).
Let us remark that 2.5 implies for n 6 3 the following property: if x1 , . . . , xn ∈ V
are linearly independent over K , then also over L. Moreover, this property holds
for every n ∈ N if and only if G is a proper subgeometry of P(W ).
APCS232.tex; 5/02/1998; 13:54; v.7; p.9
96
C.-A. FAURE AND A. FRÖLICHER
3. Quotients
We saw in 2.1 that every initial morphism is an embedding, i.e. an initial
monomorphism. However, not every final morphism is a quotient, i.e. a final
epimorphism.
EXAMPLE 3.1. Let G1 be any and G2 a non-empty discrete projective geometry
`
(i.e. G2 is a coproduct of singletons). Then the canonical map j1 : G1 → G1 G2
is a final morphism of
P roj but not an epimorphism.
`
Proof. Let g: G1 G2 − → G be a partial map into a projective geometry G
for which g◦j1 : G1 − → G is a morphism.
Since G2 is discrete, g◦j2 : G2 − → G
`
is also a morphism. Hence g: G1 G2 − → G is a morphism. Since j1 is not
surjective, it is not an epimorphism by 1.8.
2
More general examples are obtained as j1 ◦ f¯ where j1 is as before and f¯ is a
final morphism.
PROPOSITION 3.2. For a morphism f : G1 \ N → G2 of projective geometries
the following conditions are equivalent:
1◦ f is a final morphism of the category P roj ;
2◦ If E ⊆ G2 and f −1 E ∪ N is a subspace of G1 , then E is a subspace of G2 .
Proof. 1 ⇒ 2. Let f −1E ∪ N be a subspace of G1 . One chooses a projective
geometry G with a point p ∈ G and considers the map h: G2 \ E → G having
constant value p. Then the map h ◦ f : G1 \ (f −1 E ∪ N ) → G is constant and
its kernel f −1 E ∪ N is by hypothesis a subspace of G1 . This implies that h ◦ f
and hence also h is a morphism, and therefore E = Ker h is a subspace.
2 ⇒ 1. Let G be a projective geometry and h: G2 \ M → G be a map such
that h ◦ f : G1 − → G is a morphism. Then for every subspace F of G one
knows that (h ◦ f )−1 F ∪ Ker(h ◦ f ) = f −1 (h−1 F ∪ M ) ∪ N is a subspace of G1
and by hypothesis this implies that h−1 F ∪ M is a subspace of G2 . This shows
that h is a morphism.
2
COROLLARY 3.3. Let f : G′ − → G be a final morphism of`P roj . Then G1 :=
Im g and G2 := G \ Im g are subspaces of G with G = G1 G2 . Furthermore,
G2 is discrete and f factors as f = j1 ◦ f¯ where f¯ is a final epimorphism and
j1 : G1 → G is the inclusion which is a final morphism by 3.1.
COROLLARY 3.4. If G is irreducible, then every non-empty final morphism
f : G′ − → G is an epimorphism.
PROPOSITION 3.5. Every projective geometry G which contains at least two
points is a quotient of the coproduct of its lines.
APCS232.tex; 5/02/1998; 13:54; v.7; p.10
97
CATEGORICAL ASPECTS IN PROJECTIVE GEOMETRY
`
set of ∆ :=
Proof. Let ∆ be the set of all lines of G. The underlying
`
δ
is
the
disjoint
union
of
the
lines
δ
.
By
i
:
δ
→
∆
and
jδ : δ → G
δ
δ∈∆
we denote the respective
embeddings.
By
the
coproduct-property
there
exists a
`
unique morphism k:
∆ → G such that k ◦ iδ = jδ for all δ ∈ ∆. Since
every point of G lies on some line, k is an epimorphism by 1.8. Furthermore,
Proposition 1.3 implies that k is a final morphism.
2
`
DEFINITION 3.6. A morphism g: G − → G′ of projective geometries is called
reflective if for any collinear points a′ , b′ , c′ in the image Im g there exist preimages a, b, c which are collinear.
The quotients encountered in 3.5 are reflective. The same holds for the quotients
considered in 3.7 and for many other quotients. In 3.8 we give an example of a
quotient which is not reflective. We found no categorical characterization of the
reflective quotients.
PROPOSITION 3.7. Let g: G − → G′ be a morphism of projective geometries
and let us consider the following conditions:
1◦ g is a
2◦ g is a
3◦ There
that g
4◦ g is a
5◦ g is a
reflective epimorphism and satisfies (M5 );
final epimorphism and satisfies (M5 );
exists a subspace N ⊆ G and an isomorphism u: G/N → G′ such
= u ◦ π where π : G − → G/N is the canonical projection;
surjective homomorphism;
retraction of the category P roj .
One has 1 ⇔ 2 ⇔ 3 ⇔ 4 ⇒ 5; and if G is irreducible and G′ contains more
that one point, then also 4 ⇐ 5 holds.
Proof. 1 ⇒ 2. It is enough to show that every reflective epimorphism is
final, i.e. satisfies condition 2 of 3.2. So we suppose that E ⊆ G′ is such that
g−1 E ∪ Ker g is a subspace of G. Let x′ 6= y ′ ∈ E and z ′ ∈ x′ ⋆ y ′ . By 1.8 g is
surjective and so there exist x, y, z ∈ G \ Ker g with x′ = gx, y ′ = gy, z ′ = gz .
Since g is reflective we can choose x, y, z such that ℓ(x, y, z). Then one has
x 6= y ∈ g−1 E ⊆ g−1 E ∪ Ker g and hence z ∈ g−1 E ∪ Ker g. Since z ∈
/ Ker g
one has z ∈ g−1 E and hence z ′ = gz ∈ E . So we have proved that E is a
subspace of G′ .
2 ⇒ 3. We consider the canonical projection π : G \ N → G/N , where
N := Ker g. For x, y ∈ G \ N one has gx = gy ⇒ πx = πy by (M5 ), and
πx = πy ⇒ gx = gy by (M2 ). Therefore there exists a (unique) injective map
u: G/N → G′ such that g = u ◦ π . Since g is surjective, u is also surjective, i.e.
u is a bijection. One has u−1 ◦ g = π and hence the finality of g implies that u−1
is a morphism. Since π is a reflective epimorphism satisfying (M5 ) it is final by
1 ⇒ 2. Hence g = u ◦ π implies that u is a morphism.
3 ⇒ 4. We first show that the canonical projection π : G \ N → G/N is
a homomorphism. The verification of (M5 ) is trivial. We verify (M4 ). So let
APCS232.tex; 5/02/1998; 13:54; v.7; p.11
98
C.-A. FAURE AND A. FRÖLICHER
πa 6= πb and C ∈ πa ⋆ πb. Since π is surjective, C = πc for some c ∈ G \ N .
By 5.1.4 in [4] one has c ∈ (a ⋆ b) ∨ N . Hence c ∈ c′ ∨ N for some c′ ∈ a ⋆ b and
so one has C = πc = πc′ ∈ π(a ⋆ b). Knowing now that u and π are surjective
homomorphisms, the same holds for g = u ◦ π .
4 ⇒ 1. The only property of g which needs verification is the reflectivity. So
suppose that one has ℓ′ (gx, gy, gz). We may assume that gx 6= gy . Then one
gets gz ∈ gx ⋆ gy ⊆ g(x ⋆ y) and hence gz = gze for some ze ∈ x ⋆ y .
3 ⇒ 5. Since π is known to be a retraction, cf. 6.1.4 in [4], also g = u ◦ π is
a retraction.
5 ⇒ 4 is known to hold under the assumptions that G is irreducible and that
G′ contains more than one point, cf. 6.2.3 in [4].
2
EXAMPLE 3.8. A final epimorphism with empty kernel, which however is not
reflective.
We consider g := Pf : P(F52 ) − → P(F34 ) for the semilinear map f defined by
f (α, β, γ, δ, ε) := (α + βi, γ + δi, ε),
where {1, i} is a basis of F4 over F2 . We remark that i2 = 1 + i. The point of
P(K n ) represented by (α1 , α2 , . . . , αn ) ∈ K n \0 will be denoted by (α1 : α2 : . . . :
αn ). The verification that g is surjective and hence by 1.8 an epimorphism is
straightforward. Since g is not injective (e.g. (1 : 0 : 0 : 0 : 0) and (0 : 1 : 0 : 0 :
0) have the same image) and since Ker g = ∅ one concludes that (M5 ) does not
hold. Moreover, the three points
(1 : 0 : 1) = g(1 : 0 : 0 : 0 : 1)
(0 : i : 1) = g(0 : 0 : 0 : 1 : 1)
(i : 1 : 1) = g(0 : 1 : 1 : 0 : 1)
are collinear, but their unique preimages are not collinear. Therefore g is not
reflective. In order to show that g is final, let H1 := P(F42 × {0}) and H2 :=
P(F24 × {0}). For x ∈ P(F34 ) the inverse image g−1 x is a line if x ∈ H2 and a
singleton otherwise. Moreover one has g(H1 ) = H2 . Let now E ⊆ P(F34 ) be a
subset such that g−1 E is a subspace of P(F52 ). By 3.2 it is enough to show that
then E is a subspace of P(F34 ).
Case 1. card(E ∩ H2) = 0. Since H1 is a hyperplane, g−1 E contains at most
one point and hence card(E) 6 1.
Case 2. E ∩ H2 = {x} is a singleton. Then either g−1 E = g−1 x or g−1 E =
/ H1 . In the second case, using that card(C(g−1 x ∪ a)) =
C(g−1 x ∪ a) where a ∈
7, one gets card(E) = 1 + 4 = 5, and E ⊆ C(x, g(a)). Therefore E is equal to
the subspace x ⋆ g(a).
APCS232.tex; 5/02/1998; 13:54; v.7; p.12
CATEGORICAL ASPECTS IN PROJECTIVE GEOMETRY
99
Case 3. card(E ∩ H2 ) > 2. Then g−1 E contains two disjoint lines of H1 .
Hence either g−1 E = H1 or g−1 E = P(F52 ) and therefore either E = H2 or
E = P(F34 ).
PROPOSITION 3.9. Let f : V → W be a semilinear map. Then the morphism
Pf : PV − → PW is a quotient if and only if for the subsets A of W one has:
(λ · A ⊆ A for all λ ∈ L and f −1A + f −1A ⊆ f −1A) ⇒ A + A ⊆ A.
Proof. Suppose first that g := Pf is a quotient, and let A ⊆ W satisfy λ·A ⊆
A for all λ ∈ L and f −1 A + f −1 A ⊆ f −1A. Let E := {[a] ∈ P(W )/a ∈ A \ 0}
where [a] denotes the point of P(W ) represented by a. One shows that then
g−1 E ∪ N , where N := P(ker f ), is a subspace of G1 . This implies by 3.2
that E is subspace of P(V ), and from this A + A ⊆ A follows easily. For the
converse one proceeds analogously.
2
COROLLARY 3.10. Let f : V → W be a surjective semilinear map. Then
P(f ): P(V ) − → P(W ) is a final morphism of the category P roj .
Proof. Let A ⊆ W be such that f −1 A + f −1 A ⊆ f −1 A. Then one gets
A + A = f f −1A + f f −1A = f (f −1A + f −1 A) ⊆ f (f −1A) = A.
From this the assertion follows by the preceding proposition.
2
PROPOSITION 3.11. Let f : V → W be a semilinear map and dim(W ) > 2.
Then the following conditions are equivalent:
1◦ P(f ) is a retraction of P roj ;
2◦ f is surjective and quasilinear;
3◦ f is a retraction of V ec.
Proof. 1 ⇒ 2. Since P(W ) contains at least two points, the morphism g :=
P(f ) is non-constant. So one can use 2 of 1.6 in order to obtain the quasilinearity
of f (g is a homomorphism by 3.7).
2 ⇒ 3. One verifies that the induced map f : V / ker f → W is an isomorphism
of V ec and that any projection V → V /V ′ is a retraction of V ec.
3 ⇒ 1 is trivial.
4. Factorization of Morphisms
LEMMA 4.1. Let h: G1 \ N → G2 be a morphism satisfying (M5 ) and E ⊆ G1
any subspace containing N . Then one has h−1 (h(E \ N )) = E \ N .
Proof. Let a ∈ h−1 (h(E \ N )) be given. Then ha = hb for some point
b ∈ E \ N , and one may assume a 6= b. By hypothesis there exists a point
c ∈ (a ⋆ b) ∩ N . So one obtains a ∈ b ⋆ c ⊆ b ∨ N ⊆ E , and the equality
follows.
2
APCS232.tex; 5/02/1998; 13:54; v.7; p.13
100
C.-A. FAURE AND A. FRÖLICHER
PROPOSITION 4.2. Let h: G0 − → G1 be a surjective homomorphism and
g: G0 − → G any morphism. Then there exists a morphism f : G1 − → G such
that g = f ◦ h if and only if Ker h ⊆ Ker g. Moreover, the morphism f is unique
and its kernel is given by Ker f = h(Ker g \ Ker h).
Proof. We use that h is a final epimorphism satisfying (M5 ); cf. 3.7. If g =
f ◦ h one gets Ker g = (Ker h) ∪ h−1 (Ker f ) ⊇ Ker h. Conversely, suppose that
Ker h ⊆ Ker g. By the lemma one knows that Ker g = (Ker h) ∪ h−1 (N1 ), where
N1 := h(Ker g \ Ker h). Now let a, b ∈ G1 \ Ker g with a 6= b and ha = hb.
By (M5 ) there exists a point c ∈ (a ⋆ b) ∩ Ker h. So one obtains ga = gb by
(M2 ), and this shows that there exists a partial map f : G1 \ N1 → G such that
g = f ◦ h. This partial map is a morphism because h is final, and it is unique
because h is an epimorphism.
2
COROLLARY 4.3. Every morphism g: G0 − → G can be decomposed as g =
f ◦ π where π is the canonical projection G0 − → G0 / Ker g. Moreover, the
induced morphism f : G0 / Ker g → G has empty kernel.
LEMMA 4.4. Let h1 : G0 − → G1 and h2 : G0 − → G2 be two surjective homomorphisms. If Ker h1 = Ker h2 , then there exists an isomorphism ϕ: G1 → G2
such that h2 = ϕ ◦ h1 .
Proof. By 4.2 there exist a morphism ϕ: G1 − → G2 such that h2 = ϕ ◦ h1
and a morphism ψ: G2 − → G1 such that h1 = ψ ◦ h2 . From h2 = ϕ ◦ ψ ◦ h2 one
obtains ϕ ◦ ψ = id, because h2 is an epimorphism. Similarly, one has ψ ◦ ϕ = id,
and this shows that ϕ is an isomorphism.
2
DEFINITION 4.5. Let C1 , . . . , Cn be n classes of morphisms of a category X . We
say that X has (C1 , . . . , Cn )-factorization if every morphism g: G0 → Gn can be
written as g = cn ◦· · ·◦c1 with morphisms ci : Gi−1 → Gi belonging to Ci , and if
this decomposition is essentially unique: that is, if g = c′n ◦· · ·◦c′1 is another such
decomposition, then there exist isomorphisms ϕi : Gi − → G′i (i = 1, . . . , n − 1)
such that the following diagram commutes:
G0
c1
G1
/
c2
ϕ1
id
G2
/
···
Gn−1
cn
Gn
/
ϕn−1
ϕ2
id
G0
/
c′1
G′1
/
c′2
G′2
···
G′n−1
/
c′n
Gn
PROPOSITION 4.6. The category P roj has (H, K)-factorization, where H is
the class of all surjective homomorphisms and K is the class of all morphisms
having empty kernel.
Proof. The existence of a decomposition g = k ◦ h follows directly from 4.3.
Now let g = k1 ◦ h1 and g = k2 ◦ h2 be two such decompositions. We have to
show that there exists an isomorphism ϕ such that h2 = ϕ ◦ h1 and k2 ◦ ϕ = k1 .
APCS232.tex; 5/02/1998; 13:54; v.7; p.14
CATEGORICAL ASPECTS IN PROJECTIVE GEOMETRY
101
Clearly, one has Ker h1 = Ker h2 = Ker g. Therefore by 4.4 there exists an
isomorphism ϕ such that h2 = ϕ ◦ h1 . Finally, one gets k2 ◦ ϕ = k1 because
k2 ◦ ϕ ◦ h1 = k1 ◦ h1 and h1 is an epimorphism.
2
PROPOSITION 4.7. Let i: G1 − → G0 be an initial morphism and g: G − →
G0 any morphism. Then there exists a morphism f : G − → G1 such that g = i◦f
if and only if Im g ⊆ Im i. Moreover, such a morphism f is unique; its image is
given by Im f = i−1 (Im g).
Proof. We put N := Ker g. If g = i ◦ f , then Im g = i(f (G \ N )) is contained
in Im i. Conversely, suppose that Im g ⊆ Im i. Then for any point a ∈ G \ N
there exists a unique point b ∈ G1 such that ga = ib. Defining f a := b one gets
a partial map f : G \ N − → G1 with i ◦ f = g. This partial map is a morphism
because i is initial, and it is unique because i is a monomorphism, cf. 2.1. 2
COROLLARY 4.8. Every morphism g: G − → G0 can be decomposed as g =
i ◦ f where i is the inclusion of the subgeometry S(Im g) generated by Im g, and
also as g = s ◦ f where s is the inclusion of the subspace C(Im g).
DEFINITION 4.9. A morphism g: G1 − → G2 of projective geometries is called
S -dense if G2 = S(Im g). It is called C -dense if G2 = C(Im g). The morphism f
of the preceding corollary is S -dense in the first case, and C -dense in the second
one.
REMARK 4.10. It is easy to verify that the class of C -dense morphisms is
closed under composition. But we do not know whether the same holds for
the class of S -dense morphisms. This is the reason why we use a definition of
factorizations which is slightly different from the classical one; cf. Definition 15.1
and Proposition 15.5 in [2].
LEMMA 4.11. Let i1 : G1 → G0 and i2 : G2 → G0 be two initial morphisms. If
Im i1 = Im i2 , then there exists an isomorphism ϕ: G1 → G2 such that i2 ◦ϕ = i1 .
Proof. By 4.7 there exist two morphisms ϕ: G1 − → G2 and ψ: G2 − → G1
such that i2 ◦ ϕ = i1 and i1 ◦ ψ = i2 . From i2 = i2 ◦ ϕ ◦ ψ one obtains ϕ ◦ ψ = id,
because i2 is a monomorphism. Similarly one has ψ ◦ ϕ = id, and hence ϕ is an
isomorphism.
2
The proposition and its corollary give rise to two different decompositions of a
morphism g: G1 − → G2 . We first need some elementary results.
LEMMA 4.12. Let i: G1 → G2 be an initial morphism. Then i(S1 ) is a subgeometry of G2 for every subgeometry S1 ⊆ G1 , and i−1 (S2 ) is a subgeometry of
G1 for every subgeometry S2 ⊆ G2 .
Proof. By 2.1 we may assume that i is the inclusion of a subgeometry. Then
i(S1 ) = S1 and i−1 (S2 ) = G1 ∩ S2 are subgeometries.
2
APCS232.tex; 5/02/1998; 13:54; v.7; p.15
102
C.-A. FAURE AND A. FRÖLICHER
COROLLARY 4.13. Let i: G1 → G2 be an initial morphism. Then for any
subset A ⊆ G1 one has i(S(A)) = S(i(A)).
Proof. Clearly, one has S(i(A)) ⊆ i(S(A)) because i(S(A)) is subgeometry
containing i(A). Similarly, one has S(A) ⊆ i−1 (S(i(A))) because i−1 (S(i(A)))
is a subgeometry containing A. So the second inclusion follows.
2
PROPOSITION 4.14. The category P roj has (DS , I )-factorization, where DS
is the class of all S -dense morphisms and I is the class of all initial morphisms.
Proof. The existence of a decomposition g = i ◦ d follows from 4.8. Now let
g = i1 ◦ d1 and g = i2 ◦ d2 be two such decompositions. By 4.13 one has
i1 (S(Im d1 )) = S(i1 (Im d1 )) = S(Im g) = S(i2 (Im d2 )) = i2 (S(Im d2 )),
and this implies that Im i1 = Im i2 . By 4.11 there exists an isomorphism ϕ such
that i2 ◦ ϕ = i1 . Finally, one has ϕ ◦ d1 = d2 because i2 ◦ ϕ ◦ d1 = i2 ◦ d2 and i2
2
is a monomorphism. So the decomposition is essentially unique.
LEMMA 4.15. Let s: G1 → G2 be a section. Then for any subset A ⊆ G1 one
has s(C(A)) = C(s(A)).
Proof. The image s(E1 ) of a subspace E1 ⊆ G1 is a subspace of G2 because
s is a homomorphism. The inverse image s−1 (E2 ) of a subspace E2 ⊆ G2 is a
subspace of G1 because s is a morphism. So one concludes as in 4.13.
2
PROPOSITION 4.16. The category P roj has (DC , S )-factorization, where DC
is the class of all C -dense morphisms and S is the class of all sections.
Proof. One proceeds as in Proposition 4.14.
2
By combining the three Propositions 4.6, 4.14 and 4.16, we conclude that any
morphism g: G1 − → G2 can be decomposed in an essentially unique way
into four factors: a surjective homomorphism h, an S -dense morphism k having
empty kernel, a C -dense initial morphism i and a section s.
DECOMPOSITION THEOREM 4.17. The category P roj of projective geometries has (H, K ∩ DS , I ∩ DC , S )-factorization, where
•
•
•
•
H is the class of all surjective homomorphisms,
K ∩ DS is the class of all S -dense morphisms having empty kernel,
I ∩ DC is the class of all C -dense initial morphisms,
S is the class of all sections.
Moreover, one has for any morphism g: G1 − → G2 a canonical decomposition
as follows:
h
k
i
s
G1 − → G1 / Ker g −→ S(Im g) −→ C(Im g) −→ G2
APCS232.tex; 5/02/1998; 13:54; v.7; p.16
CATEGORICAL ASPECTS IN PROJECTIVE GEOMETRY
103
where h is the canonical projection and i, s are the respective inclusion morphisms.
Proof. The preceding results imply that the factorization g = sikh exists and
is of the claimed type. We now verify the uniqueness. Let g = s1 i1 k1 h1 and
g = s2 i2 k2 h2 be two such decompositions. Clearly, s ◦ i ◦ k has empty kernel.
Therefore there exists by 4.6 an isomorphism ϕ1 such that ϕ1 h1 = h2 . Now
one remarks that s ◦ i is an initial morphism and that k ◦ h is S -dense (because
Im(k ◦ h) = Im k). By 4.14 there exists an isomorphism ϕ2 such that ϕ2 k1 h1 =
k2 h2 and s2 i2 ϕ2 = s1 i1 . One has ϕ2 k1 = k2 ϕ1 because ϕ2 k1 h1 = k2 ϕ1 h1 (and
h1 is an epimorphism). Finally, one remarks that i ◦ k ◦ h is C -dense, which
follows from the equality
C(Im(i ◦ k ◦ h)) = C(Im(i ◦ k)) = CS(i(Im k))
= C(i(S(Im k))) = C(Im i).
By 4.16 there exists an isomorphism ϕ3 such that s2 ϕ3 = s1 . Using the equality
s2 ϕ3 i1 = s2i2 ϕ2 one deduces that ϕ3 i1 = i2 ϕ2 . Therefore the decomposition of
a morphism g: G1 − → G2 is essentially unique.
2
References
1. Adámek, J.: Theory of Mathematical Structures, D. Reidel Publishing Company, Dordrecht,
1983.
2. Adámek, J., Herrlich, H. and Strecker, G. E.: Abstract and Concrete Categories, Wiley, New
York, 1990.
3. Crapo, H. H. and Rota, G.-C.: On the Foundations of Combinatorial Theory: Combinatorial
Geometries, MIT Press, 1970.
4. Faure, C.-A. and Frölicher, A.: Morphisms of projective geometries and of corresponding lattices, Geom. Dedicata 47 (1993), 25–40.
5. Faure, C.-A. and Frölicher, A.: Morphisms of projective geometries and semilinear maps, Geom.
Dedicata 53 (1994), 237–262.
6. Faure, C.-A. and Frölicher, A.: Dualities for infinite-dimensional projective geometries, Geom.
Dedicata 56 (1995), 225–236.
7. Hermes, H.: Einführung in die Verbandstheorie, Grundlehren Band 73, Springer-Verlag, 1967.
APCS232.tex; 5/02/1998; 13:54; v.7; p.17