Formal Concept Analysis - Mathematical Foundations 2nd Edition
Formal Concept Analysis - Mathematical Foundations 2nd Edition
Formal Concept
Analysis
Mathematical natural stagnant
Second Edition
temporary maritime
plash, sea,
puddle lagoon
Second Edition
Bernhard Ganter Rudolf Wille (1937-2017)
Institute of Algebra Fachbereich Mathematik
Technische Universität Dresden Technische Universität Darmstadt
Dresden, Germany Darmstadt, Germany
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
v
vi Preface to the first edition
implications and dependencies between attributes are dealt with. The third chap-
ter supplies the basic notions of a structure theory for concept lattices, namely
part- and factor structures as well as tolerance relations. In each case the extent
to which these can be elaborated directly within the contexts is studied.
These mathematical tools are then used in the fourth and fifth chapter, in or-
der to describe more complex concept lattices by means of decomposition and
construction methods. Thus, the concept lattice can be split up into (possibly
overlapping) parts, but it is also possible to use the direct product of lattices or
of contexts as a decomposition principle. A further approach is that of substitu-
tion. In accordance with the same principles, it is possible to construct contexts
and concept lattices. As an additional construction principle, we shall describe a
method of doubling parts of a concept lattice.
The structural properties examined in mathematical lattice theory, for exam-
ple the distributive law and its generalizations or notions of dimension, play a
role in Formal Concept Analysis as well. This shall be treated in the sixth chap-
ter. The seventh chapter finally deals with structure-comparing maps, examining
various kinds of morphisms. Particular attention is given to the scale measures,
occuring in the context of conceptual scaling.
We limit ourselves to a concise presentation of ideas for reasons of space.
Therefore, we endeavour to give a complete reference to further results and the
respective literature at the end of each chapter. However, we have only taken into
account such contributions closely connected with the topic of the book, i.e., with
the mathematical foundations of Formal Concept Analysis. The index contains
all technical terms defined in this book, and in addition some particularly im-
portant keywords. The bibliography also serves as an author index.
The genesis of this book has been aided by the numerous lectures and activ-
ities of the “Forschungsgruppe Begriffsanalyse” (Research Group on Concept
Analysis) at Darmstadt University of Technology. It is difficult to state in detail
which kind of support was due to whom. Therefore, we can here only express
our gratitude to all people who contributed to the work presented in this book.
Two years after the German edition, this English translation has been finished.
In its content there are only a few minor changes. Although there is vivid activity
in the field, the mathematical foundations of Formal Concept Analysis have been
stable over the last years.
The authors are extremely grateful to Cornelia Franzke for her precise and
cooperative work when translating the book. They would also like to thank K.A.
Baker, P. Eklund and R.J. Cole, M.F. Janowitz, and D. Petroff for their careful
proofreading.
Preface to the second edition
For December 13, 1979, the minute book for the almost daily after lunch semi-
nar („Mittagsseminar“) of Rudolf Wille’s research group at TU Darmstadt notes
his lecture on “Lattice Theory as Algebra of Concepts”, in which he explained
how the relation between concept extents and concept intents can be understood
as a Galois connection and how therefore complete lattices can be interpreted
as algebraic structures of concept hierarchies. This was preceded by months of
lively discussion in another seminar, dealing with the meaning and significance
of mathematical order theory. Two authors had a decisive influence on this sem-
inar and thus on the genesis of the Formal Concept Analysis:
• the mathematician Garrett Birkhoff with his application-oriented view of lat-
tice theory,1 and
• the educationalist Hartmut von Hentig with his critical yet constructive under-
standing of science.2
It was in this seminar that Wille presented his definition of formal concepts for
the first time, according to K. E. Wolff [423], who reports in more detail on this
development (another source is [418]). Years of intense productivity followed,
during which Wille worked out the theory of Formal Concept Analysis with his
collaborators and students.
Sadly, Rudolf Wille, the founder and idea provider of Formal Concept Analy-
sis, passed away in January 2017. A new edition of this book had already been
planned in 2003, when Rudolf was still in good health. The book had proven it-
self, so we wanted to get by with only a few changes. These included Contextual
Attribute Logic, which later was incorporated as the first section of a new eighth
chapter. However, it took twenty years before work on a new edition began in
earnest. In the meantime, the first edition (published in German 1996 [161],
in English 1999 [163], in Chinese 2005 [164]) had been cited several thousand
times.
The variety of publications on Formal Concept Analysis, including quite a
number of books, is almost impossible to keep track of. A first, but by no means
complete, impression is given by the proceedings of the conference series ICFCA
(International Conference on Formal Concept Analysis) and CLA (Concept Lattices
and their Applications). Many authors have dealt with applications and with the
algorithms necessary for them, i.e., with topics that are only marginally men-
tioned in this book. But even if only those contributions that concern the “math-
ematical foundations” are considered, there are far too many to be included in
the book.
1 G. Birkhoff: Lattice Theory. Amer. Math. Soc. Providence, 1st edition 1940, 2nd (revised) edi-
tion 1948, 3rd (new) edition 1967
2 H. von Hentig: Magier oder Magister? Über die Einheit der Wissenschaft im Verständigungsprozeß.
vii
viii Preface to the second edition
When it became clear that the second edition of this book was only a few
weeks away from completion, I asked potential readers for a critical prelimi-
nary reading. An invitation was distributed via [email protected],
the standard mailing list for Formal Concept Analysis. To my delight, I have
been contacted by many interested readers from a wide variety of backgrounds,
and the responses that came from these contacts have noticeably improved the
manuscript. All the comments were friendly, constructive, and tried to encourage
the author, which was inspiring. Some have reported how they have benefited
from the first edition, others have suggested additional topics that should be in-
cluded or at least cited. However, this was mostly in vain, as the manuscript is
already extensive enough. And more than a few have honored the author (and
future readers) by spending a great deal of their personal time pointing out er-
rors, typos, and misleading wording.
My heartfelt thanks go to all those involved, namely Alexandre Albano, Jaume
Baixeries, Mayukh Bagchi, Radim Belohlavek, Daniel Borchmann, Dmitry Ig-
natov, René Jansen, Francesco Kriegel, Sergei O. Kuznetsov, Léonard Kwuida,
Thomas Morgenstern, Amedeo Napoli, Sergei Obiedkov, Vladimir Parkhomenko,
Uta Priss, Sebastian Rudolph, Baris Sertkaya, Gerd Stumme, Francisco Valverde-
Albacete, Karl Erich Wolff.
It is certainly no easy task to be the editor of a publishing house who over-
sees the creation of a book like this one. It takes a great deal of expertise and
sensitivity to support the often stubborn authors in their creative work while
maintaining the quality standards set by the publisher. Alexandru Ciolan, the
editor in charge at Springer, has done an exemplary job and deserves a special
word of thanks.
Finally, I thank Uta Wille for allowing her late father’s name to continue as
author. I am grateful for her trust in my work on the second edition.
Bernhard Ganter
ix
Contents
Preface
to the first edition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
to the second edition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii
0 Order-theoretic foundations 1
0.1 Ordered sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
0.2 Complete lattices . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
0.3 Closure operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
0.4 Galois connections . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
0.5 Notes, references, and trends . . . . . . . . . . . . . . . . . . . . . 20
xi
xii Contents
References 335
Index 363
Chapter 0
Order-theoretic foundations
The symbol ♦ indicates the end of a definition, > the end of an example. The “end-of-proof”
symbol is also used in this book if no proof is given.
1. x R x (reflexivity)
2. x R y and x 6= y together imply not y R x (antisymmetry)
3. x R y and y R z together imply x R z (transitivity)
For an order relation, we usually use the symbol ≤ instead of R (and ≥ instead
of R−1 ), and we write x < y for x ≤ y and x 6= y. As usual, we read x ≤ y as “x
is less than or equal to y”, etc. An ordered set is a pair (M, ≤), with M being a set
and ≤ an order relation on M .1 ♦
Examples of ordered sets are numerous in this book. The following are certainly
familiar to the reader: The real numbers R with the usual ≤-relation, but also the
space Rn with
the natural numbers N with the divisibility relation | ; the powerset P(X) of all
subsets of any set X with set inclusion. Even the equality relation = is a (trivial)
example of an order. >
d d
d d d d d d d
d
d d d d d Jd d Jd d d d d d
J J
d d d d d d d d
d Jd Jd d dJ d d d d d d d d d
J J J
Jd d d Jd d d d d @d d d
J J @
d d d d d d d d d d
d d
A A A A
d d A d A d d d d d A d A d d d d d d d
Figure 0.1 Line diagrams of all ordered sets with up to four elements.
[b, c] := {x ∈ M | b ≤ x ≤ c}.
The set c c c
(a] := {x ∈ M | x ≤ a} d
x ≤ y ⇒ ϕx ≤ ϕy
x ≤ y ⇐ ϕx ≤ ϕy,
x ≤ y ⇒ ϕx ≥ ϕy
Definition 7 The (direct) product of two ordered sets (M1 , ≤) and (M2 , ≤) is
defined to be the ordered set (M1 × M2 , ≤) with
× (M , ≤) := × M , ≤ with
t∈T
t
t∈T
t
c c c
H
c c @c c c c c @ c H c c c @ c
@ @ HH @
∼
× =
HHHHHH
c c @ cHH c
Hc
@c@@c
@ @ @ @ @ H @
H
H
HH
@ c
@
H
H
Definition 8 To be able to define the cardinal sum or disjoint union of two or-
dered sets, we first introduce the notation
Ṁt := {t} × Mt .
The sets Ṁ1 and Ṁ2 will then be disjoint copies of M1 and M2 . We define
This definition is also easily generalized in the case of any number of sum-
mands. ♦
The Duality Principle for ordered sets. The inverse relation ≥ of an order rela-
tion ≤ is also an order relation. It is called the order which is dual to ≤. A line
diagram of the dual ordered set (M, ≤)d := (M, ≥) can be obtained from the line
diagram of (M, ≤) by a horizontal reflection. If (M ≤) = ∼ (N, ≤)d , the two orders
are called dually isomorphic.
One obtains the dual statement Ad of an order-theoretic statement A (which,
apart from purely logical components, contains only the symbol ≤), by replacing
all occurences of the symbol ≤ in A by ≥. A holds for an ordered set if and only
if Ad holds for the dual ordered set. This Duality Principle is used to simplify
definitions and proofs. If a theorem asserts two statements that are dual to each
other, one usually proves only one of them, the other one follows “dually”, i.e.,
with the same proof for the dual order.
Definition 9 Let (M, ≤) be an ordered set and A a subset of M . A lower bound
of A is an element s of M with s ≤ a for all a ∈ A. An upper bound of A is
defined dually. If there is a largest element in the set ofVall lower bounds of A, it
is called the infimum of A and is denoted by inf A orW A; dually, a least upper
bound is called supremum and denoted by sup A or A. If A = {x, y}, we also
write x ∧ y for inf A and x ∨ y for sup A. Infimum and supremum are frequently
also called meet and join. ♦
If an order R on M is a subset of a second, say S, then S is called an order
extension of R. If S is even a linear order, then it is called a linear extension of
R. Each order is the intersection of its linear extensions, and the smallest num-
ber of linear extensions needed to represent the order as an intersection is its
order dimension. This dimension is discussed in detail in Section 6.4, although a
different, yet equivalent definition is used there.
2 Other names for these elements are top and bottom, other symbols are > and ⊥.
6 0 Order-theoretic foundations
d d d d
@ @
d d d @d d d
@d
@
@ @ @
d d d @d @d d @d d
@ @ @
d @d d @d @ d
from which it follows that L 6= ∅ for every complete lattice. Every non-empty
finite lattice is a complete lattice.
The order relation can be reconstructed from the lattice operations infimum
and supremum by
x ≤ y ⇔ x = x ∧ y ⇔ x ∨ y = y.
If T isWan index set and X :=V{xt | t ∈ T } aVsubset of L, instead of X we also
W
write t∈T xt and instead of X we write t∈T xt . The operations of the supre-
mum and infimum, respectively, are associative. The familiar particular case of
the associative laws, i.e., x∧(y∧z) = (x∧y)∧z, respectively x∨(y∨z) = (x∨y)∨z,
can be generalized as follows: If {Xt | t ∈ T } is a set of subsets of L, then
_ _ _ [ ^ ^ ^ [
Xt = Xt and dually Xt = Xt .
t∈T t∈T t∈T t∈T
The Duality Principle for lattices. The definitions of a lattice and a complete
lattice, respectively, are self-dual: If (L, ≤) is a (complete) lattice, then so is
(L, ≤)d = (L, ≥). Therefore, the Duality Principle for ordered sets carries over
to lattices: We obtain the dual statement of an order-theoretic statement when
we replace the symbols ≤, ∨, ∧, , , 0L , 1L etc. by ≥, ∧, ∨, , , 1L , 0L etc.
W V V W
Proposition 1 An ordered set in which the infimum exists for every subset is a complete
lattice.
Proof Let X be any subset of the ordered set. We have to prove that the supre-
mum of X exists. The set S of all upper bounds of X has an infimum s (even if
S is empty). Every element of X is a lower bound of S, i.e., is less than or equal
to s. Hence s itself is an upper bound of X and consequently the supremum.
Examples of lattices.
1) For every set M the powerset P(M ), i.e., the set of all subsets of M , is ordered
by set inclusion ⊆ and (P(M ), ⊆) is a complete lattice. In this case the lattice
operations supremum and infimum are set union and intersection.
0.2 Complete lattices 7
2) Every closed real interval [a, b] in its natural order forms a complete lattice
([a, b], ≤) with the usual infimum and supremum, respectively, as lattice opera-
tions. The ordered set (R, ≤), on the other hand, is a lattice, but it is not complete:
It lacks a greatest and a least element.
We will give further examples of complete lattices from mathematics in the
next section and throughout this book. >
Definition 11 For an element v of a complete lattice L := (L, ≤) we define
_
v∗ := {x ∈ L | x < v}
^
and v ∗ := {x ∈ L | v < x}.
3 Read: supremum-irreducible
4 Read: infimum-irreducible
8 0 Order-theoretic foundations
_
T ⊆U ⇒ T ∈ U,
(D∧ ) x ∧ (y ∨ z) = (x ∧ y) ∨ (x ∧ z)
(D∨ ) x ∨ (y ∧ z) = (x ∨ y) ∧ (x ∨ z)
♦
Distributivity and complete distributivity are self-dual properties: if they
hold for a lattice L, they also hold for Ld . The above-mentioned properties trans-
fer to complete sublattices. Powerset lattices are completely distributive.
5 ϕ(X) here stands for {ϕx | x ∈ X}.
0.3 Closure operators 9
For the special case S := {0, 1}, x0,t := x and x1,t := xt for all t ∈ T , the
property of complete distributivity yields the weaker law
_ _
(D∧ W ) x∧ xt = (x ∧ xt ).
t∈T t∈T
The dual law (D∨ V ) holds in the lattice of the closed sets of any given topolog-
ical space, and (D∧ W ) holds in the lattice of all open sets. Those lattices are not
completely distributive in general.
Remark This book deals almost exclusively with complete lattices. Unlike lat-
tices, however, they are not algebraic structures with fundamental operations of
fixed arity. The complete lattices therefore do not form an equational class in the
sense of Universal Algebra. Certain algebraic constructions, such as free algebras,
are therefore only available with restrictions, if at all.
ϕ(X ∪ Y ) = ϕ(X ∪ ϕY ).
Aϕ := {ϕX | X ⊆ G}
ϕA ϕ = ϕ as well as AϕA = A.
{A ∈ A | X ⊆ A} ⊇ {A ∈ A | Y ⊆ A}.
• X ∈ A ⇔ X = {A ∈ A | X ⊆ A} ⇔ ϕA X = X ⇔ X ∈ AϕA .
T
since ϕX ∈ Aϕ .
Every closure system A can be understood as the set of all closures of a closure
operator. Therefore, the terms “closure” and “closed set” are used as synonyms.
0.3 Closure operators 11
(2) subgroups: For any group G, the set U(G) of all subgroups of G is a clo-
sure system. The complete lattice (U(G), ⊆) is called subgroup lattice of G.
Provided that G is commutative, U1 ∨ U2 = U1 + U2 , and more generally,
_
X = {u1 + u2 + . . . + un | there are U1 , . . . , Un ∈ X
with ui ∈ Ui for i ∈ {1, . . . , n}}.
(3) closed sets: For a topological space T (for example, for Rn ), the set A(T )
of all closed sets of T is a closure system. In the complete lattice (A(T ), ⊆)
the supremum is equivalent to the topological closure of the union, i.e.,
_ [
X= X.
12 0 Order-theoretic foundations
(4) convex sets: For Rn the set C(Rn ) of all convex subsets is a closure system,
i.e., (C(Rn ), ⊆) is a complete lattice and in this lattice the supremum is the
convex closure of the union.
(5) the faces of polyhedra: For a polyhedron P , the set S(P ) of all faces of P is
a closure system. The complete lattice (S(P ), ⊆) is called the face lattice of
P ; for those lattices there is no general “good” description of the suprema.
(6) equivalence relations: For a set M , the set E(M ) of all equivalence rela-
tions on M is a closure system on M × M . The complete lattice (E(M ), ⊆)
is called the lattice of equivalence relations of M ; in this lattice
_
X = {(a, b) ∈ M × M | there are R1 , . . . , Rn ∈ X and
(xi , xi+1 ) ∈ Ri for i ∈ {1, . . . , n}
with a = x1 and b = xn+1 }.
>
Theorem 1 can be used to determine the system of all closures of a given closure
operator. However, that method is cumbersome. For practical computations we
introduce the Next Closure algorithm, see Algorithm 1. It assumes that the
base set G is finite and linearly ordered (in an arbitrary way). The closures are
created one after the other. The previously generated closure is input for the
generation of the next one. It starts with the closure of the empty set, i.e. with ϕ∅.
The closures are generated in lectic order, defined as follows: A subset A ⊆ G
is lectically smaller than a subset B ⊆ G if A 6= B and the smallest element in
which A and B differ belongs to B.
0.3 Closure operators 13
More about the Next Closure algorithm will be said in Section 2.1, notably in
Theorem 5 (p. 81). The refinements presented in the Theorems 6 and 60 (p. 285)
generally work for closure systems.
Example To demonstrate how the algorithm works, we apply it to the closure
system of all closure systems on a fixed base set {1, . . . , m}, see Algorithm 2.
Each such closure system is a subset F ⊆ P({1, . . . , m}) of the powerset of
{1, . . . , m}. For the linear order, as required by Algorithm 1, we choose one which
extends the ⊇-order, i.e., with A ⊇ B ⇒ A ≤ B. A simple example of such an
order is obtained by encoding the subsets of {1, . . . , m} by the bit representations
of the integers 0, . . . , 2m −1 and then using the reverse order 0 > 1 > . . . > 2m −1.
Moreover, the closure system must be equipped with its generator map L : F → F
defined as follows: L[A] is the smallest set in F such that
\
A = {B ∈ F | B ≤ L[A], A ⊆ B}.
Theoretically, this algorithm can be (and has been) used for counting: just
start with the trivial 1-element closure system and repeat. However, the number
of closure systems grows so quickly that a complete listing is out of reach already
for m = 7, see [80]. >
14 0 Order-theoretic foundations
Generators
An explicit formula is obtained from this using Möbius inversion (see Stan-
ley [349]): X
numgen(C) = 2|D| µ(D, C),
D⊆C,D=ϕD
Proof Let S ⊆ T and assume that S is not a minimal generator. Then there is
some proper subset of S, say R, such that ϕR = ϕS. With Proposition 3 it follows
that ϕT = ϕ((T \ S) ∪ S) = ϕ((T \ S) ∪ ϕS) = ϕ((T \ S) ∪ ϕR) = ϕ((T \ S) ∪ R).
Thus (T \ S) ∪ R is a proper subset of T generating the same closure as T . So T
is no minimal generator either.
Example Consider two closure systems on {a, b, c},
Some generalized finiteness condition is needed to ensure that each closed set
has a minimal generator, such as the condition of chain finiteness, which requires
that each chain of closed sets is finite, see p. 41. If each closed set has a minimal
7It has been proved to be a #P -complete problem (Kuznetsov [244]).
8“Minimal” in the sense of “non-redundant”, but not necessarily of “minimum size”. A closed
set may have several minimal generators of different cardinalities.
0.4 Galois connections 15
generator, then the number of closed sets is obviously less than or equal to the
number of minimal generators. This is exploited in Section 6.1 for the proof of
an upper bound on the number of formal concepts. The (chain finite) closure
systems, where each closed set has a unique minimal generator, are characterized
in Theorem 52 on p. 265.
Proposition 5 states that the set of minimal generators of a closure system on
G is always an order ideal9 in the powerset (P(G), ⊆) and thus a concept extent
of the free distributive lattice FCD(G), see Construction (4) on p. 48. Conversely,
for finite G each non-empty order ideal of (P(G), ⊆) is the set of minimal gen-
erators of a closure system (take as closed sets G together with all non-maximal
elements of the order ideal). This shows that the above example does not present
an exceptional case: There are 61 closure systems on {a, b, c} (see Figure 7.4 on
p. 291), but only 19 non-empty order ideals in (P({a, b, c}), ⊆).
C ∈ A ⇔ α(C) ∈ B.
This notion is more restrictive than that of lattice isomorphism: Each closure sys-
tem, when ordered by set inclusion ⊆, forms a complete lattice, and isomorphic
closure systems have isomorphic lattices. But the converse is not true. It hap-
pens that non-isomorphic closure systems give isomorphic lattices. This will be
explored in more detail when it comes to clarifying and reducing formal contexts
in Section 1.2.
In addition to isomorphisms, there are other useful mappings to compare clo-
sure systems. A mapping α : G → H is called closure continuous if the preimage
of every closed set ist closed, formally, if α−1 (C) ∈ A for every C ∈ B. The scale
measures, which play a central role in Chapter 7 of this book, are closure contin-
uous maps.
x ≤1 ψy ⇔ y ≤2 ϕx. ♦
If this condition is met, then both mappings ϕ and ψ are order-reversing and their
compositions ψ ◦ ϕ and ϕ ◦ ψ are closure operators on L1 and L2 , respectively,
with dually isomorphic lattices of closed sets (comp. Propositions 6, 7, and 10
below).
Remark The name “Galois connection” comes from Galois theory, a branch of
algebra named after the mathematician Évariste Galois (1811–1832). There an
extension field E (satisfying certain conditions) of a field F is examined on the
basis of its Galois group, which consists of those automorphisms of E that leave F
pointwise fixed. The lattice L1 then is the lattice of intermediate fields, i.e., subfields
of E containing F . L2 is the subgroup lattice of the Galois group. ϕ maps an
intermediate field to its pointwise stabilizer, ψ maps a subgroup to the set of its
fixed points, which turns out to be an intermediate field. (The corresponding
context construction can be found on page 58).
In Galois theory a simplified notation is common, where the same symbol is
used for the two mappings ϕ and ψ, namely a simple “prime”, as it is also used
for the derivative of a function.
In fact, this Galois connection can be traced back to a single relation between
the field E and the Galois group G, namely the relation “is fixed by”. For an
intermediate field H and a subgroup S of G one has
H 0 = {α ∈ G | αh = h for all h ∈ H}
S 0 = {e ∈ E | αe = e for all α ∈ S}.
The condition in Definition 17 uses neither infima nor suprema and can there-
fore be generalized to arbitrary ordered sets without modification. An even fur-
ther generalization will be discussed in Section 3.4.
Definition 18 A pair (ϕ, ψ) of maps ϕ : P → Q and ψ : Q → P between two
ordered sets (P, ≤1 ) and (Q, ≤2 ) is called a Galois connection if it satisfies the
Galois condition p ≤1 ψq ⇔ q ≤2 ϕp for all p ∈ P and all q ∈ Q. The two maps
then are called dually adjoint to each other.11 ♦
See Figure 0.4 for an example.
Proposition 6 A pair (ϕ, ψ) of maps is a Galois connection if and only if all three of the
following conditions hold:
1. p1 ≤ p2 ⇒ ϕp1 ≥ ϕp2
2. q1 ≤ q2 ⇒ ψq1 ≥ ψq2
3. p ≤ ψϕp and q ≤ ϕψq .
Proof p ≤ ψq yields ϕp ≥ ϕψq because of 1., and from this we get ϕp ≥ q using
3., that is, one direction of the Galois condition. The other follows symmetrically.
11 In the following, we mostly use the same symbol ≤ for both ≤1 and ≤2 .
0.4 Galois connections 17
Figure 0.4 A pair of dually adjoint maps between two ordered sets.
Proposition 9 A map
ϕ : L1 −→ L2
between complete lattices has a dual adjoint if and only if
_ ^
ϕ xt = ϕxt
t∈T t∈T
holds for xt ∈ L1 .
Proof If ψ is dually adjoint to ϕ, then by the Galois condition
^
y≤ ϕxt ⇔ y ≤ ϕxt for all t ∈ T
t∈T
⇔ xt ≤ ψy for all t ∈ T
_
⇔ xt ≤ ψy
t∈T
_
⇔y≤ϕ xt .
t∈T
Since we have not presupposed that M and N are disjoint, this notation allows
ambiguous formulations, which, however, can easily be avoided.
Proof The Galois condition is easily verified. It implies that (ϕR , ψR ) is a Galois
connection and that the two sets used in order to define R(ϕ,ψ) are equal. Ac-
cording to this definition (x, y) ∈ R(ϕ,ψ) ⇔ y ∈ ϕ{x} and thus, by Proposition 9,
\
ϕX = ϕ{x}
x∈X
\
= ϕR(ϕ,ψ) {x}
x∈X
= ϕR(ϕ,ψ) X,
i.e., ϕR(ϕ,ψ) = ϕ and correspondingly ψR(ϕ,ψ) = ψ. The last statement R(ϕR ,ψR ) =
R follows immediately from the equivalence x ∈ ψR {y} ⇔ xRy.
The use of the term “Galois connection” is not uniform. Some authors prefer
to replace one of the ordered sets by its dual. We prefer to call such pairs of maps
adjunctions. In the case of complete lattices we obtain:
ϕ(x) ≤ y ⇔ x ≤ ψ(y).
In this case, ϕ is called a residuated map, ψ is called the residual map, and the maps
are adjoint to each other. If one of the maps is injective, the other one is surjective, and
vice versa.
Proof This is an immediate consequence of Proposition 9 and the preceding
propositions: If we replace L2 by the dual lattice Ld2 , ϕ and ψ form a Galois con-
nection. The relation between injectivity and surjectivity can be inferred from
Proposition 7.
Rudolph [324, 325] uses the lower and upper bounds given by Burosch et al. [71]
to investigate the possibilities of succintly representing closure systems. A dia-
gram of the closure systems on a three-element set is shown in Figure 7.4 on
p. 291.
Kuznetsov [247] (see also the earlier papers cited there) introduces a notion
of stability of closed sets based on the relative number of their generators. This
is closely related to the Möbius function. Hermann & Sertkaya [189] study com-
plexity issues connected with minimal generators.
Chapter 1
Concept lattices of formal contexts
The basic notions of Formal Concept Analysis are those of a formal context and a
formal concept. The adjective “formal” is meant to emphasize that we are dealing
with mathematical notions, which only reflect some aspects of the meaning of
context and concept in standard language. However, an overly frequent mention
of the suffix “formal” occasionally disrupts the flow of language. Thus, it shall
be understood that where we write context or concept we actually mean a formal
context or a formal concept, respectively.
a b c d e f g h i
1 | leech × × ×
2 | bream × × × ×
3 | frog × × × × ×
4 | dog × × × × ×
5 | waterweed × × × ×
6 | reed × × × × ×
7 | bean × × × ×
8 | maize × × × ×
Figure 1.1 Formal context of an educational film “Living Beings and Water”. The at-
tributes are: a: needs water to live, b: lives in water, c: lives on land, d: needs chlorophyll to
produce food, e: two seed leaves, f: one seed leaf, g: can move around, h: has limbs, i: suckles its
offspring.
A0 := {m ∈ M | g I m for all g ∈ A}
(the set of attributes common to the objects in A). Correspondingly, for a set B
of attributes we define
B 0 := {g ∈ G | g I m for all m ∈ B}
The proposition shows that the two derivation operators form a Galois con-
nection between the powerset lattices P(G) and P(M ) (see Section 0.4). Hence
we obtain (by Proposition 10) two closure systems on G and M , which are dually
order-isomorphic to each other:
For every set A ⊆ G, A0 is an intent of some concept, since (A00 , A0 ) is always a
concept. A00 is the smallest extent containing A. Consequently, a set A ⊆ G is an
extent if and only if A = A00 . The same applies to intents. The union of extents
generally does not result in an extent. On the other hand, the intersection of any
number of extents (respectively intents) is always an extent (intent), as is proved
by the following proposition:
Proposition 13 If T is an index set and, for every t ∈ T , At ⊆ G is a set of objects,
then
[ 0 \
At = A0t .
t∈T t∈T
way is denoted by B(G, M, I) and is called the concept lattice of the formal con-
text (G, M, I). ♦
Example The formal context in Figure 1.1 has 19 concepts. The line diagram in
Figure 1.2 represents the concept lattice of this formal context. >
123
45678
ag ac ab ad
1234 5678
34 12
agh 678 356 adf
234 568
abg acd
acgh abc abdf
34 123 678 56
36
abgh acdf
23 68
acghi abcgh acde abcdf
4 3 7 6
abcdef
ghi
Figure 1.2 Concept lattice for the formal context of Figure 1.1. A much better way of la-
beling is shown in Figure 1.4.
Proof of the Basic Theorem. First, we will explain the formula for the infimum.
Since At = Bt0 for each t ∈ T ,
1.1 Formal context and formal concept 27
\ [ 00
At , Bt
t∈T t∈T
i.e., it has the form (X 0 , X 00 ) and is therefore certainly a concept. That this can
only be the infimum, i.e., the largest common subconcept of the concepts (At , Bt ),
follows immediately from the fact that the extent of this concept is exactly the
intersection of the extents of (At , Bt ). The formula for the supremum is sub-
stantiated correspondingly. Thus, we have proven that B(G, M, I) is a complete
lattice.
Now we prove, first for the special case (L, ≤) = B(G, M, I), the existence of
mappings γ̃ and µ̃ with the required properties. We set
holds for every formal concept (A, B), i.e., γ̃(G) is supremum-dense and µ̃(M )
is infimum-dense in B(G, M, I). More generally, if ϕ : B(G, M, I) → (L, ≤) is an
isomorphism, we define γ̃ and µ̃ by
γ̃ : G → L, µ̃ : M → L
h ∈ {g ∈ G | γ̃g ≤ x} ⇔ γ̃h ≤ x
⇔ γ̃h ≤ µ̃n for all n ∈ {m ∈ M | x ≤ µ̃m}
⇔ h I n for all n ∈ {m ∈ M | x ≤ µ̃m}
⇔ h ∈ {m ∈ M | x ≤ µ̃m}0 .
B(M, G, I −1 ) ∼
= B(G, M, I)d ,
and
(B, A) 7→ (A, B)
is an isomorphism.
In other words: if we exchange the roles of objects and attributes, we obtain
the dual concept lattice. Thus, the Duality Principle extends to concept lattices.
The mappings γ̃ and µ̃ which appear in the Basic Theorem indicate how the
context can be identified in the concept lattice. This is further deepened by the
following definition.
Definition 24 For an object g ∈ G we write g 0 instead of {g}0 for the object intent
{m ∈ M | g I m} of the object g. Correspondingly, m0 := {g ∈ G | g I m} is
the attribute extent of the attribute m. Retaining the symbols used in the Basic
1.2 Formal context and concept lattice 29
Theorem, we write γg for the object concept (g 00 , g 0 ) and µm for the attribute
concept (m0 , m00 ). ♦
The line diagram in Figure 1.2 shows the intent and the extent of every formal
concept. The labelling can be simplified considerably by entering each object and
each attribute only once, namely to the corresponding object or attribute concept
(see Figure 1.3). One can then still read the formal context as well as all concept
extents and concept intents from the line diagram: If you look for the concept
extent of a formal concept represented by a small circle, this consists of the ob-
jects located at this circle or at the circles which can be reached by descending
line paths from this circle. Correspondingly, the intent can be found by follow-
ing all line paths going upwards from the circle and noting down the attributes
assigned to these circles.
g c b d
h f
1 5
2 8
i e
4 3 7 6
The sparing, reduced labelling enables us to enter the full names of the objects
and attributes of the formal context in Figure 1.1 into the diagram. This improves
the readability of the diagram, as can be seen in Figure 1.4.
A formal context (G, M, I) can be easily reconstructed from the system of all
its formal concepts. G and M appear as the extent and the intent of the triv-
30 1 Concept lattices of formal contexts
leech waterweed
Figure 1.4 Concept lattice for the educational film “Living beings and water”.
ial boundary concepts: The set of all objects is the extent of the largest concept,
(∅0 , ∅00 ) = (G, G0 ). Dually, M is the intent of the least concept, (∅00 , ∅0 ) = (M 0 , M ).
The incidence relation I is given by
[
I = {A × B | (A, B) ∈ B(G, M, I)}.
It is even easier to read off the context from the concept lattice, as the Basic The-
orem shows. On the other hand, concept lattices of different contexts can well
be isomorphic to each other. The context manipulations which do not alter the
structure of the concept lattice include the merging of objects with the same in-
tents and attributes with the same extents, respectively:
Definition 25 A formal context (G, M, I) is called clarified, if for any objects
g, h ∈ G it always follows from g 0 = h0 that g = h and, correspondingly, m0 = n0
implies m = n for all m, n ∈ M . ♦
Example Figure 1.5 shows a context representing the service offers by an office
supply shop. Below is the clarified context. >
Also not affecting the structure of the concept lattice are attributes that can
be written as combinations of other attributes. More precisely: If m ∈ M is an
attribute and X ⊆ M is a set of attributes with m 6∈ X but m0 = X 0 , then the
attribute concept µm is the infimum of the attribute concepts µx, x ∈ X, i.e., the
set µ(M \ {m}) is also infimum-dense in B(G, M, I), and according to the Basic
Theorem
B(G, M, I) ∼= B(G, M \ {m}, I ∩ (G × (M \ {m}))).
1.2 Formal context and concept lattice 31
Computers
e
Training,
Workshops @
@ Copy machines,
@ typewriters
@@e
Service contracts
Furniture e
Planning J
J eSpecialized machines
J Instruction
J
J
e
JJ
Consulting,
Assembly and installation,
Original spare parts and accessories,
Repairs
Figure 1.6 The concept lattice for the context of Figure 1.5.
This is called the standard context of the lattice L. For practical work with for-
mal contexts, the proposition has the following consequences: Every finite for-
mal context can be brought into a reduced form without changing the structure
of the concept lattice, and the latter is unique. We first clarify the context, i.e., we
merge objects with the same intents and attributes with the same extents. Then
we delete all objects, the intent of which can be represented as the intersection
of other object intents, and correspondingly all attributes, the extent of which is
the intersection of other attribute extents.
It is easy to reconstruct the concepts of the original formal context from those
of the reduced context, if one has kept a record of the reduction process. If
4See also Proposition 16.c).
5Two formal contexts (G1 , M1 , I1 ) and (G2 , M2 , I2 ) are called isomorphic if there are bijective
mappings α : G1 → G2 , β : M1 → M2 with g I1 m ⇔ (αg) I2 (βm) for all g ∈ G1 , m ∈ M1 , see
Definition 97 on p. 283.
1.2 Formal context and concept lattice 33
c
c c × × c
c × c × c @c × c
c c × @c c
c c c
× × c @c × × c × × c @c
× @c × × c @c × × c @c
× c @c × @c
c c
c @c
× × c c@ c × × × c
c c @c
@
× × c@ c@ c × ×
\ c
× × @c × @ @ c × \c
c c
c @@c c @
× × × × @c
× × c@
@c c × × c@
@c
× × @c
@ × @c
@
Figure 1.7 Reduced formal contexts with up to three objects, and their concept lattices.
The context (∅, ∅, ∅) was omitted, as was its (one-element) concept lattice.
we denote, for a finite clarified context (G, M, I), the set of its irreducible ob-
jects by Girr and the set of irreducible attributes by Mirr , the reduced context is
(Girr , Mirr , I ∩ (Girr × Mirr )), and each concept (A, B) of (G, M, I) corresponds to
the formal concept (A ∩ Girr , B ∩ Mirr ) of (Girr , Mirr , I ∩ (Girr × Mirr )). For every
object g ∈ G and every extent A of (G, M, I)
g ∈ A ⇔ g 00 ∩ Girr ⊆ A ∩ Girr ,
holds dually for the attributes. If we note down the set g 00 ∩ Girr for every re-
ducible object g and the set m00 ∩ Mirr for every reducible attribute m, it is easy
to obtain the concepts of (G, M, I) from those of (Girr , Mirr , I ∩ (Girr × Mirr )).
There is another way to carry out the reduction of a clarified context, namely
with the help of arrow relations, which we will define next and which will be-
come important for other reasons later on. These relations can conveniently be
entered into the cross table, since they only apply to “empty cells”, i.e., to object-
attribute pairs which do not stand in the relation I.
34 1 Concept lattices of formal contexts
Non-Aligned
Non-Aligned
Group 77
Group 77
OACPS
OACPS
OPEC
OPEC
SIDS
SIDS
LDC
LDC
Afghanistan × × × East Timor × × × ×
Algeria × × × Ecuador × ×
Angola × × × × × Egypt × ×
Antigua and Barbuda × × × × El Salvador ×
Argentina × Equatorial Guinea × × × ×
Armenia Eritrea × × × ×
Azerbaijan × × Eswatini × × ×
Bahamas × × × × Ethiopia × × × ×
Bahrain × × Fiji × × × ×
Bangladesh × × × Gabon × × × ×
Barbados × × × × Gambia × × × ×
Belize × × × × Ghana × × ×
Benin × × × × Grenada × × × ×
Bhutan × × × Guam
Bolivia × × Guatemala × ×
Botswana × × × Guinea × × × ×
Brazil × Guinea-Bissau × × × × ×
Brunei × × Guyana × × × ×
Burkina Faso × × × × Haiti × × × × ×
Burundi × × × × Honduras × ×
Cambodia × × × India × ×
Cameroon × × × Indonesia × ×
Cape Verde × × × × Iran × × ×
Central African Republic × × × × Iraq × × ×
Chad × × × × Ivory Coast × × ×
Chile × × Jamaica × × × ×
China × Jordan × ×
Colombia × × Kazakhstan
Comoros × × × × × Kenya × × ×
DR Congo × × × × Kiribati × × × ×
Rep. Congo × × × × Kuwait × × ×
Cook Islands × × Kyrgyzstan
Costa Rica × Laos × × ×
Cuba × × × × Lebanon × ×
Djibouti × × × × Lesotho × × × ×
Dominica × × × × Liberia × × × ×
Dominican Republic × × × × Libya × × ×
Figure 1.8 Membership of developing countries in supranational groups. (Part 1). LDC =
Least Developed Countries, SIDS = Small Island Developing Countries, OPEC = Petrol Ex-
porting Countries, OACPS = African, Caribbean and Pacific States.
1.2 Formal context and concept lattice 35
Non-Aligned
Non-Aligned
Group 77
Group 77
OACPS
OACPS
OPEC
OPEC
SIDS
SIDS
LDC
LDC
Madagascar × × × × Samoa × × ×
Malawi × × × × São Tomé + Príncipe × × × × ×
Malaysia × × Saudi Arabia × × ×
Maldives × × × × Senegal × × × ×
Mali × × × × Seychelles × × × ×
Marshall Islands × × × Sierra Leone × × × ×
Mauritania × × × × Singapore × × ×
Mauritius × × × × Solomon Islands × × × ×
Micronesia × × × Somalia × × × ×
Moldova South Africa × × ×
Mongolia × × South Sudan × × ×
Morocco × × Sri Lanka × ×
Mozambique × × × × Sudan × × × ×
Myanmar × × × Suriname × × × ×
Namibia × × × Syria × ×
Nauru × × Tajikistan ×
Nepal × × × Tanzania × × × ×
Nicaragua × × Thailand × ×
Niger × × × × Togo × × × ×
Nigeria × × × × Tonga × × ×
Niue × × Trinidad + Tobago × × × ×
North Korea × × Tunisia × ×
North Macedonia Turkmenistan × ×
Oman × × Tuvalu × × ×
Pakistan × × Uganda × × × ×
Palau × × Un. Arab Emirates × × ×
Palestine × × Uruguay ×
Panama × × Uzbekistan ×
Papua New Guinea × × × × Vanuatu × × × ×
Paraguay × Venezuela × × ×
Peru × × Vietnam × ×
Philippines × × Yemen × × ×
Qatar × × Zambia × × × ×
Rwanda × × × × Zimbabwe × × ×
Saint Kitts and Nevis × × × ×
Saint Lucia × × × ×
St. Vincent + Grenadines × × × ×
Figure 1.8 Membership of developing countries in supranational groups. (Part 2). For data
sources, see Section 1.5.
36 1 Concept lattices of formal contexts
OPEC
Figure 1.9 Concept lattice of the formal context of developing countries, shown in Figure 1.8.
1.2 Formal context and concept lattice 37
/ I and
(g, m) ∈
g . m :⇔
if g 0 ⊆ h0 and g 0 6= h0 , then h I m,
/ I and
(g, m) ∈
g % m :⇔
if m0 ⊆ n0 and m0 6= n0 , then g I n,
g%. m : ⇔ g . m and g % m.
♦
Thus, g . m if and only if g 0 is maximal among all object intents which do
not contain m. In other words: g . m holds if and only if g does not have the
attribute m, but m is contained in the intent of every proper subconcept of γg. If
we now let _
(γg)∗ := {x ∈ B(G, M, I) | x < γg}
as in Definition 11, then (γg)∗ is a subconcept of γg and γg is -irreducible, if
W
and only if γg 6= (γg)∗ . This, on the other hand, is equivalent to the fact that there
is an attribute m in the intent of (γg)∗ which is not contained in the intent of γg,
i.e., to g . m for some m ∈ M . Therefore, we obtain
g . m ⇔ γg ∧ µm = (γg)∗ 6= γg
g % m ⇔ γg ∨ µm = (µm)∗ 6= µm.
Example Figure 1.10 shows the context from Figure 1.5 with the arrow relations;
beside it the reduced context.
× × × ×
× × .
% . × .
%.
.
% × × × % × ×
.
% × % % × %
.
% × × .
%
Figure 1.10 Context with arrow relations, and the reduced context.
The significance of the arrow relations for the reduction of a formal context is
shown by the next proposition:
Furthermore, the following statements hold for every finite6 formal context:
c) γg is V -irreducible ⇔ There is an m ∈ M with g %
W
. m.
d) µm is -irreducible ⇔ There is a g ∈ G with g % . m.
A complete lattice (L, ≤) is called doubly founded, if for any two elements x < y
of L there are elements s, t ∈ L with:
s is minimal with respect to s ≤ y, s 6≤ x, as well as
t is maximal with respect to t ≥ x, t 6≥ y.
♦
By means of Proposition 15 we realize easily that the attribute n and the object
h that appear in Definition 28 must be irreducible. The same applies to the lattice
elements sVand t in the second part of the definition: s must be -irreducible and
W
t must be -irreducible. This means that the property “doubly founded” implies
the existence of “many” irreducible elements.
Proposition 16
a) Every finite formal context is doubly founded.
b) A formal context which does neither contain infinite chains g1 , g2 , . . . of objects
with g10 ⊂ g20 ⊂ . . . nor infinite chains m1 , m2 , . . . of attributes with m01 ⊂ m20 ⊂
. . . is doubly founded.
c) Each formal concept of a doublyV founded context is the supremum of -irreducible
W
concepts and the infimum of - irreducible concepts. Hence Proposition 14 also ap-
plies to concept lattices of doubly founded contexts.
d) If (G, M, I) is doubly founded and g ∈ G, m ∈ M , the following hold true: if
g . m, then there is an attribute n with g %. n, and if g % m, then there is an object
h with h % . m. Hence parts c) and d) of Proposition 15 also apply to doubly founded
contexts.
e
@@
J(L) supremum-dense
@ eM (L) infimum-dense
@
e
coalgebraic @ algebraic
e @ A@
@ e
supremum-founded @ @ einfimum-founded
@ @
e @e
@
@ @
@ @
e @ @ e A@ e e
@ @ @ A@
@
@ @ @
atomistic @ @ ecoatomistic
e @e e @@ e
@
@ A@
@ @ @
@ @ @ @
e @ @ e A@ e @ @ e@ e Ae
@ @ @ A@
@ @
@ @ df @ @
@@ @e A@
@ e @ @ e
@ @A@
@e @ @ e
@
B(N, PE (N), ∈)@ @ @ @ @ B(N, PE (N), ∈)d
@ e A@ e
@ @ @ e@
@ A@
@ e @ @ e@
@ Ae
@ @ @ @
A@ e @ @ e e @ @ e A e
@ @ @A@
@ @ @
e@
@ @ @ @
B(C[0,1] ) @@ @ e@ cf @ @ B(C[0,1] )d
A@
@e @ e Ae
@ @ f e@ @@ A
A@ e @ @ e@ @@ e
@ @ @
@ @ @
e@@ @ e@ A e@@ e
e@ A
@ @ @e
B(I3 )
@ @ e J B( I 3 ) d
@
@
B(S) @@ e
J
B( S ) d
@ eJ
B(N, N, 6=) J B(N, N, =)
it follows that A = A00 . An ordered set is chain finite, if every chain contained
by it is finite.
The lattice presented in Figure 1.11 is the result of an attribute exploration in
accordance with Section 2.3, i.e., the represented implications between the prop-
erties are really provable. We omit the proofs.
Formal contexts are very simple mathematical objects. It is not surprising that
they are easy to construct and that there is an almost inexhaustible variety of
constructions. Some of them are collected in this section. We start with simple
and frequently used composition methods, with sums, products and unions of
formal contexts, etc. We usually formulate only the composition of two formal
contexts, but mostly the definitions can be generalized to arbitrary many. The
additional statements on the concept lattices of the resulting contexts carry over.
This part is followed by context constructions from mathematical structures, es-
pecially ordered sets. Our list is diverse, but far from complete. Further construc-
tions are described in the later chapters of this book, but of course also in the
literature.
What is presented here are, for the time being, purely mathematical construc-
tions. Their interpretation and their meaning for conceptual knowledge process-
ing is not yet addressed here. The subsequent section gives first hints.
Some of the context compositions require the parts to be disjoint. We will use
the abbreviations Ġj := {j} × Gj , Ṁj := {j} × Mj and I˙j := {((j, g), (j, m)) |
(g, m) ∈ Ij } for j ∈ {1, 2} in the following definition.
Definition 29 Let K := (G, M, I) and Kt := (Gt , Mt , It ) for i ∈ {1, 2} be formal
contexts. It is:
Kc := (G, M, (G × M ) \ I) the complementary context to K,
Kd := (M, G, I −1 ) the dual context to K,
and, if G = G1 = G2 ,
K1 K2 := (G, Ṁ1 ∪ Ṁ2 , I˙1 ∪ I˙2 ) the apposition of K1 and K2 ,
as well as dually, if M = M1 = M2 ,
K1
:= (Ġ1 ∪ Ġ2 , M, I˙1 ∪ I˙2 ) the subposition of K1 and K2 .
K2
Kcd is called the contrary context to K,
K | Kc is the dichotomization of K, and
K1 ∪˙ K2 := (Ġ1 ∪ Ġ2 , Ṁ1 ∪ Ṁ2 , I˙1 ∪ I˙2 ) the disjoint union of K1 and K2 .
♦
42 1 Concept lattices of formal contexts
Remark: By using Ġi for {i} × Gi and Ṁi , respectively, we intend to make sure
that the sets are disjoint. However, strictly speaking, apposition and subposition
under this definition become non-associative. We will overlook this fact and tac-
itly identify the contexts
( K1 K2 ) K3 and K1 ( K2 K3 ) .
The same applies to the subposition, even to hybrid forms of the two operations.
We do not distinguish between
K1 | K2 K1 K2
and .
K3 | K4 K3 K4
× := (G, M, G × M )
∅ := (G, M, ∅)
are occasionally used without further describing the sets G and M , if they are
evident from the context. For example
K1 ×
∅ K2
denotes the context (Ġ1 ∪ Ġ2 , Ṁ1 ∪ Ṁ2 , I˙1 ∪ I˙2 ∪ (Ġ1 × Ṁ2 )), the concept lattice of
which is isomorphic to the vertical sum of the concept lattices B(K1 ) and B(K2 )
(provided that K1 does not contain a full column and K2 does not contain a full
row, cf. Section 4.3).
Each extent of K1 ∪˙ K2 , apart from the extent Ġ1 ∪ Ġ2 , is entirely contained
in one of the sets Ġi . The corresponding applies to the intents. Therefore, the
concept lattice L := B(K1 ∪˙ K2 ) is a horizontal sum, i.e., it is the union L =
(L1 , ≤) ∪ (L2 , ≤) of two sublattices which only overlap in the smallest and the
largest element: L1 ∩ L2 = {0L , 1L }. Provided that there are no full rows or
columns in K1 and K2 , we have (Li , ≤) ∼ = B(Ki ) or, more generally, (Li , ≤) =
˙
B(Ġ1 ∪ Ġ2 , Ṁ1 ∪ Ṁ2 , Ii ).
(Ġ1 ∪ Ġ2 , Ṁ1 ∪ Ṁ2 , I˙1 ∪ I˙2 ∪ (Ġ1 × Ṁ2 ) ∪ (Ġ2 × Ṁ1 )). Context sum
♦
7 For the “dot” notation see the remarks before Definition 29. A more general definition is given
in Section 5.1.
1.3 Context constructions 43
The concept lattice of a context sum is isomorphic to the product of its concept
lattices. In the case of two contexts we therefore obtain
B(K1 + K2 ) ∼
= B(K1 ) × B(K2 ),
with
(g1 , g2 )∇(j, m) : ⇔ gj Ij m for j ∈ {1, 2}. ♦
× ×
× `
× ∼ × ×
a
= →
× × ××
× ×
× ×
× ×
×
× `
× ××
← × ∼
=
a
× ××
××
× ×
×××
The extents of the semiproduct are precisely the sets of the form A1 × A2 , each
set Aj `being an extent of Kj . This also yields the structure of the concept lattice
B(K1 a K2 ): Essentially, the concept lattice is the product of the concept lattices
of the factor contexts, though there is a modification regarding the zero elements.
Precisely, the instruction for the construction reads as follows: Provided the the
extent of the zero element of B(Kj ) is empty, remove that zero element from
B(Kj ) (for j ∈ {1, 2}). Then take the direct product of these ordered sets and,
if an element was removed earlier, add a new zero element to make a complete
lattice. This lattice is then isomorphic to the concept lattice of the semiproduct.
Figure 1.12 shows a small example.
44 1 Concept lattices of formal contexts
K1 × K2 := (G1 × G2 , M1 × M2 , ∇)
with
(g1 , g2 ) ∇ (m1 , m2 ) : ⇔ g1 I1 m1 or g2 I2 m2 . ♦
The concept lattice of the direct product is called the tensor product of the con-
cept lattices of the factor contexts. The tensor product will be discussed in more
detail later (Sections 4.4, 5.4). The cross table of the direct product is obtained by
replacing each empty cell in the table of K1 by a copy of K2 and each cross by a
rectangle full of crosses of the size of K2 . For an example see Figures 4.19 (page
198) and 4.20. An easy conclusion is that
One instance of the direct product deserves special mention, namely the case
where K2 is the standard context of the three-element chain, i.e.,
×
K2 = .
K ×
,
K K
and it can be shown that the concept lattice of this is isomorphic to the order
relation of B(K) (which, understood as an ordered set of pairs with component-
wise order, is itself a complete lattice). Compare the context recursion for the
free completely distributive lattices below on page 49.
Definition 33 Forming the Boolean matrix product of two contexts assumes
that the attribute set of the first is equal to the object set of the second. Let
(G, F, IGF ) and (F, M, IF M ) be such formal contexts. The Boolean matrix prod-
uct then is
(G, F, IGF ) · (F, M, IF M ) := (G, M, I)
with
(g, m) ∈ I : ⇔ ∃f ∈F (g, f ) ∈ IGF and (f, m) ∈ IF M .
♦
This product is seen less as a construction method. Rather, the task is to factorize
a given formal context K, i.e., to find an (ideally small) set F and relations IGF
and IF M to form a factorization of K. We will go into this in more detail in the
last part of Section 6.4.
Another context construction, the so-called substitution sum, where a context
is inserted into another context, will be described in Section 4.3. The sum and
1.3 Context constructions 45
the product of reduced contexts are reduced (cf. Corollary 3, p. 200). Reducible
objects or attributes with empty intents or extents may occur in the case of the dis-
joint union. Semiproducts of reduced contexts are reduced if the factors (allow-
ing for one exception at most) are atomistic, i.e., if they satisfy g 0 ⊆ h0 ⇒ g = h.
It is easy to state numerous simple arithmetical rules for context constructions,
which are useful for some proofs. In particular, the direct product is (up to iso-
morphism) commutative and associative; it is distributive over the direct sum,
the apposition and the subposition. We note down one of these results for later:
Proposition 18 We have
Proof We may assume that the three contexts Ki =: (Gi , Mi , Ii ), i ∈ {1, 2, 3},
have disjoint object sets and disjoint attribute sets. By
and
(M1 ∪ M2 ) × M3 = (M1 × M3 ) ∪ (M2 × M3 ),
the two contexts of the proposition have the same objects and attributes. For the
incidence we find the same on both sides as well, namely
g ∈ G1 and m ∈ M2 or
g ∈ G2 and m ∈ M1 or
(g, h) I (m, n) ⇔ h I3 n or
g ∈ G1 , m ∈ M1 , and g I1 m or
g ∈ G2 , m ∈ M2 , and g I2 m.
We now present some interesting context families, but must explain in advance
why we often use the synonym “scale” instead of “formal context”. The next sec-
tion explains how Formal Concept Analysis deals with data that are not given in
the form of formal contexts. The method used for this is called Conceptual Scaling.
We call formal contexts that are particularly suitable for this purpose conceptual
scales, see Definition 35 (p. 60). The simplest variants, which we call “elemen-
tary standard scales”, will be discussed in more detail in Definition 37. Besides,
the formal contexts in this list serve as a reservoir of examples for mathematical
reasoning.
First, contexts are described that can be obtained naturally from arbitrary or-
dered sets ((1)–(7)). The examples that follow come from graphs, equivalence
46 1 Concept lattices of formal contexts
relations, and other simple standard mathematical notions. Finally, some con-
structions from advanced mathematics are briefly mentioned.
(1) For every set S the contranominal scale
is reduced. The concepts of this context are precisely the pairs (A, S \ A)
for A ⊆ S. The concept lattice is isomorphic to the powerset lattice of S,
and thus has 2|S| elements. For S = {1, 2, . . . , n} it is listed as elementary
standard scales (Def. 37) under the name Nnc .
(2) From an arbitrary ordered set (P, ≤) we obtain the general ordinal scale
Its concepts are precisely the pairs (X, Y ) with X, Y ⊆ P where X is the
set of all lower bounds of Y and Y is the set of all upper bounds of X.
This concept lattice is called the Dedekind-MacNeille completion of the
ordered set (P, ≤). It is the smallest complete lattice in which (P, ≤) can
be order-embedded, in the sense of the following theorem:
e
ce ce
@ @
e e
Q e ce e c
e
Q @ @ @ @
e e
Q e
Q @ @
@ @
@e
@
c
@@e
c
@@e
Figure 1.13 Example of an ordered set (P, ≤) and its completion B(P, P, ≤).
Proof Evidently, the concepts of (P, P, ≤) are precisely the pairs (A, B) with
A, B ⊆ P and
in particular, all pairs ((x] , [x)) with x ∈ P are concepts of (P, P, ≤), which con-
firms ι as an embedding. If the supremum of X exists in (P, ≤), then
h_ \
X = [x) ,
x∈X
i.e., ι
W
X=
_ i h_ \ ↓ \ _ _
= X , X = [x) , [x) = ((x] , [x)) = ιX.
x∈X x∈X
The equation for existing infima is shown dually. With respect to the missing
part of the proof we refer to Proposition 39 (p. 119).
We mention two familiar instances of this construction.
• What are the formal concepts of O(Q,≤) = (Q, Q, ≤), where (Q, ≤) is the set of
rational numbers in their natural order? Apart from the boundary concepts
(∅, Q) and (Q, ∅), there is exactly one concept for each real number r ∈ R,
namely
((r], [r)),
where
(r] := {q ∈ Q | q ≤ r} and [r) := {q ∈ Q | r ≤ q}.
These Dedekind cuts were used by R. Dedekind to construct the real numbers
from the rational ones. Indeed,
B(O(Q,≤) ) ∼
= (R ∪ {−∞, ∞}, ≤).
which is called the contraordinal scale for (P, ≤). In this case, the formal
concepts are precisely the pairs (X, Y ) with the following properties:
• X ∪ Y = P and X ∩ Y = ∅,
48 1 Concept lattices of formal contexts
d
a b c d e f d @d
e d df a × × × × × @d
b × × × × ×
c dZ dd −→ d @d
Z
c × × × −→
d × × × @d
a dZ db
Z
e ×
f × d @d
(P, ≤) @d
(P, P, 6≥)
B(P, P, 6≥)
Figure 1.14 An ordered set (P, ≤), the corresponding contraordinal scale and its concept
lattice, i.e., the ideal lattice of (P, ≤).
An ×
A0 = ∅ and An+1 = .
An An
×
Note that An+1 = An × , as in Definition 32.
The construction can be generalized by taking an ordered set (S, ≤) as
the base set, the set OI(S, ≤) of the order ideals of (S, ≤) as the object set
and the set OF(S, ≤) of the order filters of (S, ≤) as the attribute set. The
concept lattice
is called the free completely distributive lattice over the ordered set
(S, ≤).
(5) For an arbitrary ordered set (P, ≤), we define a filter to be a subset of P
which is an order filter and in which furthermore any two elements have a
common lower bound. Hence F ⊆ P is a filter if and only if the following
two conditions are satisfied:
1. From x ∈ F and y ≥ x it follows that y ∈ F ,
2. for any two elements x, y ∈ F there is an u ∈ F with u ≤ x and u ≤ y.
Dually, an ideal is defined to be a subset of P which is an order ideal and
contains a common upper bound for any two elements contained in it.
Filters in this sense are among other things the principal filters. Dually,
each principal ideal is an ideal. The set of all filters is denoted by F (P, ≤),
the set of all ideals by I(P, ≤). We obtain the doubly founded context
where again
F ∆ I : ⇔ F ∩ I 6= ∅.
(6) Again from an ordered set (P, ≤) we obtain the general interordinal scale
the concept system of which we explain by means of the extents: the at-
tribute extents are precisely the principal ideals and the principal filters
of (P, ≤), the concept extents are all intersections of those sets. These in-
50 1 Concept lattices of formal contexts
Figure 1.15 A nested line diagram of the free distributive lattice FCD(4). Such diagrams are
introduced in 2.2. The one shown here is due to S. Thiele [369]. The method that led to it is
explained in Example 17.
1.3 Context constructions 51
clude all intervals8 of (P, ≤). In general, these are all sets which constitute
intersections of intervals.
(7) By analogy with (6) we obtain the convex-ordinal scale
In this case, the extents are precisely the convex subsets of (P, ≤), i.e.,
those subsets which contain with any two elements a and b all elements
c with a ≤ c ≤ b.
1,a 1,b 1,c 1,d 1,e 1,f 2,a 2,b 2,c 2,d 2,e 2,f
a × × × × × ×
b × × × × × ×
c × × × × × ×
d × × × × × ×
e × × × × × ×
f × × × × × ×
Figure 1.16 The convex-ordinal scale of the ordered set from Figure 1.14.
If the ordered sets in the above definitions are composable, for example as
a cardinal sum or as a direct product, then it is to be expected that the formal
contexts defined from them also can be split up. This is true, even if in different
ways, as exemplified by the following rules:
Proposition 19 We have
8 in the sense of Definition 5 (p. 3), i.e., only the “closed” intervals.
52 1 Concept lattices of formal contexts
a f
b e
c d
Figure 1.17 The concept lattice of the convex-ordinal scale from Figure 1.16.
(A, B) 7→ (B, A)
If the relation R is irreflexive, the extent and the intent of each concept
must be disjoint and we have
{x, y} X :⇔ |{x, y} ∩ X| 6= 1
|S|
we obtain a context (G, M, ) with objects and 2|S|−1 attributes,
2
which is reduced except for one full column. Every extent of this context
is a set of two-element subsets of S, i.e., it can be understood as a sym-
metric reflexive relation on S; actually, the relations occurring are pre-
cisely the equivalence relations on S. Hence the concept lattice B(G, M, )
is isomorphic to the lattice E(S) of equivalence relations. We can give a
mnemonic rule for this context series as well. We get P1 := (∅, {∗}, ∅) and
obtain the n+1-st context of this series, Pn+1 , from the n-th as follows: We
form the apposition of Pn with the cross table Prev n , which is identical to
Pn , apart from the fact that the columns are written down in the reversed
order.
Pn Prev
n
2n − 1 . . . 2n−1 2n−1 − 1 . . . 0
We add n further rows, which we fill with crosses such that the columns
of this subcontext look like the binary representations of the numbers
2n − 1, . . . , 0. An example is given in Figure 1.18.
54 1 Concept lattices of formal contexts
× ×× ×
×× ××
10011001
× × × ×
32100123 −→
××××
76543210
×× ××
× × × ×
Figure 1.18 Context P4 for the lattice of equivalence relations on a 4-element set.
Figure 1.19 shows the lattice for n = 8. Brylawski also proved that these
lattices admit a polarity (in the sense defined in (8) above), given by a 7→
a∗ with
a∗i := |{j | aj ≥ i}|, i ∈ {1, . . . , n}.
It can be deduced that there are precisely
(n + 1) · (n + 2)
−1
6
(q + 1, . . . , q + 1, q, . . . , q , 1, . . . , 1),
| {z } | {z } | {z }
r m−r s
where
n − s = q · m + r, s ≥ 0, m > r ≥ 0, q ≥ 2, and q ≥ 3 if s 6= 0 6= r.
7+1
6+2
6+1+1 5+3
5+1+1 4+4
4+3+1
5+1+1+1
4+2+2
4+2+1+1 3+3+2
3+3+1+1
4+1+1+1+1
3+2+2+1
3+2+1+1+1 2+2+2+2
3+1+1+1+1+1 2+2+2+1+1
2+2+1+1+1+1
2+1+1+1+1+1+1
1+1+1+1+1+1+1+1
K0 := L0 := × and
Ln+1 := ∅ Ln , Kn+1 :=
Kn Kn
,
Ln Ln Kn Ln
then we obtain
Σn ∼
= B(Kn ).
The contexts Kn are reduced except for the full rows and full columns. Σ4
is presented in Figure 1.20.
56 1 Concept lattices of formal contexts
e4321
HH
e3421 HH
H
HH HH
e e3241 HH e
HH
4231
HH 4312
H
HH e
e2431 e3214
HH 3412
e e e e
HH
4213 2341
4132
H 3142
H HH
H H
e e e He
H
HH 2314
2413
1432
H
3124
H
e H e e
H
2143 HH 4123
1342
e e e
H
2134
HH
HH H
H 1423 1324
H HH e
1243
HH
H
HH
H e
1234
where
X |= (A, b) : ⇔ (A ⊆ X ⇒ b ∈ X).
The extents of this contexts are precisely the closure systems on S, while
the intents, when read as sets of implications, describe the corresponding
closure operators. The arrow relations are also easy to specify, since one
gets
1.3 Context constructions 57
X . (A, b) ⇔ A ⊆ X, b ∈
/ X,
X % (A, b) ⇔ X %
. (A, b) ⇔ X = A.
It will be shown in Chapter 6 of this book that many structural properties
of these lattices can immediately be deduced from this characterization
of the arrows.
××××××××
× × × ×
×× ××
× ×× × K(d,2) K(d,2)
×××× K(d+1,2) = .
× × × × K(d,2) Kc(d,2)
×× ××
× × ××
Figure 1.21 K(3,2) := (GF(2)3 , GF(2)3 , ⊥), a formal context derived from the 3-dimen-
sional vector space over the two-element field. The help lines are intended to illustrate
the recursive construction given on the right. The first row and column are reducible. The
remaining 7 × 7-table gives the point-line incidence of the projective plane of order 2.
(H, H, ⊥)
(G, G,
b I), where g I χ : ⇔ χ(g) = 1.
The extents of the formal concepts are the closed subgroups, the intents
are their annihilators.
(17) Galois theory Let E be a field, F be a subfield and let AutF (E) denote
the group of those automorphisms of E which pointwise fix F. Consider
the formal context
The concept extents are intermediate fields between F and E the concept
intents are subgroups of AutF (E), but not necessarily all of them. It re-
quires additional conditions (“finite dimensional Galois extension”) to
make sure that every intermediate subfield is a concept extent and every
subgroup is an intent.
(18) Define for a group G the formal context
9 Our presentation follows Kerkhoff et al. [219], where precise definitions can be found, but
also a reference to the quotation.
1.4 Conceptual scaling of many-valued contexts 59
In standard language the word “attribute” is not only used for properties which
an object may or may not have. Attributes such as “color”, “weight”, “sex”,
“grade” have values. We call them many-valued attributes, in contrast to the one-
valued attributes considered so far. Data with many-valued attributes is usually
found in a form vaguely described as a data table. This is cast here into a mathe-
matical definition under the name many-valued context.
Definition 34 A many-valued context (G, M, W, I) consists of sets G, M and W
and a ternary relation I between G, M and W (i.e., I ⊆ G × M × W ) for which
it holds that
10Further information on the role of the “empty cells” in a context will be given in the notes at
the end of the chapter.
60 1 Concept lattices of formal contexts
Example The many-valued context represented in the upper part of Figure 1.23
shows a comparison of the different possibilities of arranging the engine and the
drive chain of a motorcar (cf. Figure 1.22). >
How can we assign concepts to a many-valued context? We do this in the fol-
lowing way: The many-valued context is transformed into a one-valued one, in
accordance with certain rules, which will be explained below. The concepts of
this derived one-valued context are then interpreted as the concepts of the many-
valued context. However, this interpretive process, called conceptual scaling, is
not at all unique. The concept system of a many-valued context depends on the
scaling, and scaling is an interpretive decision. This may at first be confusing, but
has proved to be an excellent instrument for a purposeful evaluation of data.
In the process of scaling, first of all each attribute of a many-valued context is
interpreted by means of a formal context. This context is called a conceptual scale.
Definition 35 A scale for the attribute m of a many-valued context is a (one-
valued) context Sm := (Gm , Mm , Im ) with m(G) ⊆ Gm . The objects of a scale
are called scale values, the attributes are called scale attributes. ♦
Every formal context can be used as a scale. Formally there is no difference
between a scale and a context. However, we will use the term “scale” only for
contexts which have a clear conceptual structure and which bear meaning. Some
particularly simple contexts are used as scales time and again. A summary (in
tabular form) of the most important ones can be found in Figure 1.28 at the end
of this section.
As already mentioned, the choice of the scale for the attribute m is not math-
ematically compelling, it is a matter of interpretation. The same is true for the
second step in the process of scaling, the joining together of the scales to make
a one-valued context. In the simplest case, this can be achieved by putting to-
gether the individual scales without connecting them. This is described below
as plain scaling. Particularly when dealing with numerical scales this may well be
unsatisfactory. In this case we need the scaling by means of a composition operator.
In the case of plain scaling the derived one-valued context is obtained from
the many-valued context (G, M, W, I) and the scale contexts Sm , m ∈ M as fol-
lows: The object set G remains unchanged, every many-valued attribute m is
replaced by the scale attributes of the scale Sm . If we imagine a many-valued
context as represented by a table, we can visualize plain scaling as follows: Ev-
ery attribute value m(g) is replaced by the row of the scale context Sm which
belongs to m(g). A detailed description will be given in the following definition,
for which we first introduce an abbreviation: The attribute set of the derived con-
text is the disjoint union of the attribute sets of the scales involved. In order to
make sure that the sets are disjoint, we replace the attribute set of the scale Sm
by
Ṁm := {m} × Mm .
as in Definition 29.
11 Source: Schlag nach! 100 000 Tatsachen aus allen Wissensgebieten. BI-Verlag 1982.
Drive Drive Road Self Space Construc- Maintain-
efficiency efficiency holding steering economy ability
empty loaded tion cost
Conventional poor good good understeering good medium excellent
Front-wheel good poor excellent understeering excellent very low good
Rear-wheel excellent excellent very poor oversteering poor low very poor
Mid-engine excellent excellent good neutral very poor low very poor
All-wheel excellent excellent good underst./neut. good high poor
Rear-wheel × × × × × × × × ×
Mid-engine × × × × × × × × × × ×
All-wheel × × × × × × × × ×
Figure 1.23 A many-valued context: Drive concepts for motorcars. Below a derived one-valued context.
61
62 1 Concept lattices of formal contexts
and
g J (m, n) : ⇔ m(g) = w and w Im n. ♦
f
@
@
f
Dl+ @
@ @
@ @
fM+@
@
@ @
R+ f @ @ fDe+
aa @ @ ! !
aa f
aa @ ! @! @
! @
Dl++ @ fCl
! @
E+ f f aa@f!! @ @f
De
@ ++
a a
@ aaaaaa
@
@ @ !!@ !! @!!!
aa aa ! !@ !! @
M-
@ @
!@
f @f aa f a f!
a !
f!
! @f @f E-
aa Su aa
@@
! ! !!
@ aa aa @ Cvl @ @! !!
@ aa Ch a a @ R++@ ! !!!@ !
@f aa f @ R- -,So
E++ fDl- fE- -,M- - ! f
M++,Cm,De- aa @ !@ ! ! !
aa Su/n
Standard aa @ Sn !!
! Rear-wheel
aa @ !AA
All-wheel Mid-engine
!
a@f
a !!
Front-wheel
Example We obtain the one-valued context in Figure 1.23 as the derived context
of the many-valued context presented above it, if we use the following scales.
The abbreviations are: De := drive efficiency when empty, Dl := drive efficiency
when loaded, R := road holding/handling properties, S := self steering effect,
E := economy of space, C := cost of construction, M := maintainability.
++ + − ++ + −−
SDe := SDl := ++ × × SR := ++ × ×
+ × + ×
− × −− ×
1.4 Conceptual scaling of many-valued contexts 63
u o n u/n ++ + − −− vl l m h
u × ++ × × vl × ×
SS := o × SE := SM := + × SC := l ×
n × − × m ×
u/n × −− × × h ×
The concept lattice is shown in Figure 1.24. If we had used the scale SE for the
attributes De, Dl and R as well, the derived context would have only turned out
slightly different. >
The mathematical definition of formal contexts allows to turn relations orig-
inating from arbitrary domains into formal contexts and to investigate their
concept lattices, i.e., even those where an interpretation of the sets G and M as
“objects” or “attributes” seems artificial. This is the case for many formal con-
texts from mathematics, whose concept lattices then have structural properties
that are highly rare in empirical data. Nevertheless, such formal contexts are of
great importance for data analysis. They are suitable, for example, as ideal struc-
tures or as scales for the conceptual scaling of many-valued contexts, introduced
above.
Standard scales
The simplest and most commonly used scales are now introduced in standard-
ized form. The elementary ones are discussed in Definition 37, a more general
selection is compiled in Figure 1.28. That these scales are standardized means the
following: The definition of scaling requires that the value set of a many-valued
attribute is a subset of the object set of the scale used. For the standardized scales
we frequently use n := {1, 2, . . . , n} as the object set. It may then be necessary to
rename the objects before scaling.
The dichotomic scale D := ({0, 1}, {0, 1}, =). The dichotomic 0 1
scale constitutes a special case, since it is isomorphic to the scales 0 ×
N2 and M1,1 and closely related to I2 . It is frequently used to scale 1 ×
attributes with values of the kind {yes, no}.
A special case of plain scaling which frequently occurs is the case that all
many-valued attributes can be interpreted with respect to the same scale or fam-
ily of scales. Thus we speak of a nominally scaled context, if all scales Sm are
nominal scales etc. We call a many-valued context nominal, if the nature of the
1.4 Conceptual scaling of many-valued contexts 65
c ≤4 ≥1
c c≤ 4
c ≤4 4
J
c 4 J c≥ 5
J ≤ 3 c @ c≥ 2 c≤ 3
≤3c 5
B@ 3 ≤ 2 c @ c @ c≥ 3 3
≤2c c≤ 2
c1 c2 c3 c4
BB @
@ 2 c≥ 6 ≤ 1 c @ c @ c @ c≥ 4 2
≤1c 6 1 2 3 4
@ B 1 Q
Q A c≤ 1
@c 1
@@B c
B @ QQA c
A
Figure 1.25 The concept lattices of the elementary scales are named after the scales. The
figure shows a nominal lattice, B(N4 ), a biordinal lattice, B(M4,2 ), an interordinal lat-
tice, B(I4 ), and an ordinal lattice, B(O4 ). The ordinal lattice B(On ) is isomorphic to the
n-element chain Cn .
Figure 1.26 Example of an ordinal context: Ratings of monuments on the Forum Ro-
manum in different travel guides. The context becomes ordinal through the number of
stars awarded.
e
AQ
Q
A QQ
Polyglott:∗ M:∗
A
A eGB:∗ Q Q eBaedecker:∗
Q
e e
Curia Temple Basilica of
of Maxentius
!! @
Basilica Julia !
!
Romulus
@
House of the Vestals @
M:∗∗ @
e e @e
Temple of
@ @
Vespasian @ @
@ @
e @@ e @ eColumn
@
Temple of Saturn of Phocas
@ @
@ @
Portico of the
@ @
GB:∗∗ @
Twelve Gods
@@ e e @e
@ @
@ @
@ @
Temple of Vesta
e @
@ e @e
@
Arch of Titus
M:
@
∗∗∗
Arch of @,
, e
Septimus Severus @
PP
Temple of Antoninus
@e
@ and Faustina
Temple of Castor
and Pollux
Figure 1.27 The concept lattice of the ordinal context from Figure 1.26. M := Michelin,
GB := Les Guides Bleus.
g J n : ⇔ An ⊆ g I .
Ocd
P (P, P, 6≥) contraordinal scale hierarchy and
independence
CP Ocd c
P OP convex-ordinal convex ordering
scale
1.1
Formal Concept Analysis has been developed from the end of the 1970s at the
Faculty of Mathematics of Darmstadt University of Technology. The first pro-
grammatic publication on Formal Concept Analysis was
68 1 Concept lattices of formal contexts
1.2
There is a simple way to assign a clarified context to every context K := (G, M, I):
12DIN stands for “Deutsche Industrienorm” and is characteristic for standards issued by the
German National Bureau of Standards.
1.5 Notes, references, and trends 69
the symbols having the following meanings: ker γ is the equivalence relation on
G with
(g, h) ∈ ker γ : ⇔ γg = γh.
ker µ is defined correspondingly. The equivalence classes of ker γ are the objects
of K◦ , those of ker µ the attributes. The incidence is defined by
The number of reduced contexts with four objects is 126. It is 13596 for five
objects, and 108096891 for six. The latter value was found by Klaus Hoffmann,
see sequence A047684 in https://oeis.org/.
Even without the additional
|G|\|M | 1 2 3 4 5 6
condition “reduced”, it is not
1 2 3 4 5 6 7
easy to determine the numbers.
2 3 7 13 22 34 50
The figures in the adjacent table
3 4 13 36 87 190 386
were calculated by M. Wiesend
4 5 22 87 317 1053 3250
in the Bayreuth group. Our
5 6 34 190 1053 5624 28576
source are lecture notes by
6 7 50 386 3250 28578 251610
A. Kerber and W. Lex [218].
7 8 70 734 9343 136758 2141733
Further values can again be
8 9 95 1324 25207 613894 17256831
found in https://oeis.org/,
see e.g., sequences A002623, A002727, and A006148.
Neither is it easy to determine the maximal possible number f (n) of attributes
in a reduced context with n objects. For small n we obtain
n 1 2 3 4 5 6
,
f (n) 1 2 4 7 13 24
where f (6) = 24 was recently found by Ignatov [197]. The question was men-
tioned on StackExchange as #3241237, and a constructive lower bound was given
there by Peter Taylor,
X i
f (n) ≥ i .
0≤i<n 2
The arrow relations have been introduced in [390] following the example
of the weak perspectivities in congruence theory (cf. [175]). There were numer-
ous forerunners. For example Day [86] already used the double arrow relation
(“relation ρ”) in order to characterize semidistributivity as well as a “relation
C”, which is closely related to the arrow relations, in order to describe the con-
gruences of finite lattices. Doubly founded lattices have first been mentioned in
[396]. Geyer [169] has examined possible configurations of the arrow relations.
Algebraic lattices in Formal Concept Analysis were studied by Hitzler et al. [193],
see also [236].
The data for Figure 1.8 were compiled from the following sources:
List of developing countries: DAC list 2023 of the German government,
LDC, SIDS: https://www.un.org/ohrlls/,
OPEC: https://www.opec.org/opec_web/en/about_us/25.htm,
OACPS: https://www.oacps.org/.
1.3
Context constructions are treated in many papers, including [134]. The comple-
mentation has been comprehensively examined in Deiters [91], [90], but see also
Berry & Sigayret [46]. Instead of (g, m) ∈
/ I we sometimes write g r
I m, as in the
following simple, but useful proposition:
Proposition 20 (Krötzsch [236]) g ∈ hII ⇔ h ∈ gr
IrI
.
Weinheimer [384] introduces the product apposition as a further construction.
For Xia’s context product see Definition 98 on p. 284. Doerfel [99] proves that
products of doubly founded complete lattices are doubly founded. The concept
lattices of the semiproduct powers of the dichotomic scale are precisely the “full
concept lattices” in Lex [262].
There is another definition of a product of contexts that suggests itself,
with
(g1 , g2 ) & (m1 , m2 ) : ⇔ g1 I1 m1 and g2 I2 m2 .
This has been considered by various authors (Schaffert [332], Reuter [315], Erné
[118]), but does not have the importance of the direct product in mathematical
literature. The extents of K1 & K2 are besides G1 × G2 precisely the sets U1 × U2
with
Ext(Ki ) if Gi is an attribute intent of Ki ,
Ui ∈
Ext(Ki ) \ {Gi } if not.
The concept lattice is therefore closely related to those of the context sum and
the semiproduct.
1.5 Notes, references, and trends 71
1.4
Research developments
The question of whether and how Formal Concept Analysis can be generalized
was also intensively pursued. This seems obvious, since the theory starts with
an extremely simple form of data, the formal context, and one can investigate
whether the same approach will lead to success with other, more sophisticated
data. However, it should not be overlooked that our approach is intentionally
simple and yet universal. The Basic Theorem (Theorem 3) states that every com-
plete lattice is isomorphic to a concept lattice. Any “generalization” that ulti-
mately leads to a complete lattice can therefore also be formulated with the help
of formal contexts. Nevertheless, such alternative paths to the same goal can be
very valuable because they provide a different language framework that may be
more intuitive or practical for certain applications.
1.5 Notes, references, and trends 73
It should also be recalled that the reason for the success of Formal Con-
cept Analysis is that it provides mathematics which supports human conceptual
thinking. Mathematical methods, even very abstract ones, are welcome if they
have the potential to facilitate human-oriented knowledge processing. Mere gen-
eralization without reference to real meaning is not our aim.
The most important addendum to the model presented in Sections 1.1 and 1.2
of this book in our view is the inclusion of many-valued contexts by means of
conceptual scaling as introduced in Section 1.4.
Lehmann and Wille [257] have outlined a triadic concept analysis, where the
incidence relation is ternary and the concepts consist of three sets. An algorithm
for finding the triadic concepts was given in [235]. Several studies have shown
that this approach can be used for applications, see e.g. Jäschke et al. [202], Be-
lohlavek & Vychodil [41], Glodeanu [172], Missaoui & Kwuida [283], Ignatov
et al [199], Rudolph et al. [326], or Felde & Stumme [127]. The mathematical
theory, however, seems to be demanding. The “Basic Theorem” for the obtained
trilattices was proved in [409] and was later generalized to the n-adic case by
Voutsadakis [382]. Biedermann [48] has presented a first systematic study. But
many fundamental questions have so far remained unanswered. Remarkable re-
cent results by Bazin and his coauthors [29, 30, 31] suggest that the answers may
be different than thought.
Fuzzy concepts
Early on, the question of how to combine Formal Concept Analysis with aspects
of Fuzzy Logic was discussed. There are related elements in the work of Pawlak
[297] and Burusco Juandeaburre & Fuentes-Gonzales [204], and Umbreit [371]
has presented a first comprehensive study. Wolff [422] draws a comparison to
conceptual scaling.
The topic was then taken up by Belohlavek, Vychodil and their collaborators
at Palacky University, Olomouc. Their careful research resulted in a large num-
ber of publications which cannot be adequately discussed here, see [34, 42] for
example, but also, for a broader setting, [33]. Much of the work of this research
group is documented in the Proceedings of the CLA conference series, available
at cla.inf.upol.cz. The “Basic Theorem” proved in [34] was later simplified by
Krupka [240].
74 1 Concept lattices of formal contexts
Faulty data
The question of how Formal Concept Analysis deals with unreliable data was
also repeatedly raised. Again, it is worth recalling the fundamental aim of this
theory development, which is to strengthen cognitive autonomy. This requires
that simplifying assumptions are made transparent, including their impact on
the results. By no means should an opaque mathematical formalism hide arti-
facts that cannot be justified rationally. It would be presumptuous to claim that
Formal Concept Analysis can simply redeem these claims. But they remain es-
sential.
It may first be necessary to specify what exactly is meant. Various authors have
dealt with incomplete (e.g., Burmeister and Holzer [70], Dubois and Prade [107],
Krupka and Lastovicka [241], [154], Zhi and Li [427]), faulty (e.g., Pensa and
Boulicaut [299], Cellier et al. [74], Borchmann [61]), and with probabilistic data
(e.g., Vityaeva et al. [375], Kriegel [230]). It seems unclear whether these ap-
proaches can interact and be unified.
Fuzzy concept analysis was already metioned as one possible approach. Some
authors have also explored the links to rough set theory, see Kent [214] for an
early contribution, Poelmans et al. [301] for a survey, and Meschke [281] for a
theoretical analysis. The Darmstadt group has focused on “formal contexts with
question marks” to handle incomplete data.
to be able to use the syntax of logical expressions. The Generalized Formal Concept
Analysis by Chaudron and Maille [75] is also similar.
There are at least two elaborated theories that are structurally closely related to
Formal Concept Analysis but have distinctly different application goals and use
different terminology. One is Barwise and Seligman’s theory of distributed sys-
tems, and is documented in [28]. The other is the theory of knowledge spaces
(and later of learning spaces) developed by Doignon and Falmagne [101, 122].
The latter provides a theory of human learning with extensive practical appli-
cations. One difference of the mathematical theory is that the knowledge states
(which roughly correspond to concept extents), are not necessarily closed under
intersections, but under set unions.
Not so close, but still related, are approaches by Diday [93] and Marty [277].
A similar attempt at restructuring with respect to mathematical logic is made by
[411].
Additional structure
How do you calculate the concept lattice and a good diagram for a given for-
mal context? In principle, this is simple and can be done by hand. We introduce
this in the first section of this chapter. However, the work becomes tedious very
quickly, and computer assistance is desired. Since the introduction of Formal
Concept Analysis, so many algorithms and implementations have emerged that
we cannot discuss them here and instead refer to [154]. These computer pro-
grams are impressively fast. A concept lattice with a billion elements can easily
be computed, and diagrams are also provided. However, computer-generated
lattice diagrams are rarely satisfactory. This is not so much because of the pro-
gramming as because it is not really understood what actually makes a good
diagram. Nested line diagrams, which we introduce alongside, allow readable
representations of systems consisting of a few hundred concepts. However, di-
agrams of larger concept lattices are usually no longer readable. For this pur-
pose, we will develop techniques in later chapters that allow a decomposition
into readable parts.
Another area in which there has been extensive development is the theory of
implications between attributes, in particular the knowledge acquisition method
of attribute exploration. This is also extensively represented in the already cited
book [154], so that we can limit ourselves here to the basic idea.
In principle, it is not difficult to find all the concepts of a formal context. The
following proposition summarizes some naive methods for doing so:
Proposition 21 Each formal concept of a formal context (G, M, I) has the form (X 00 , X 0 )
for some subset X ⊆ G and also the form (Y 0 , Y 00 ) for some subset Y ⊆ M . Conversely,
all such pairs are concepts.
Every concept extent is the intersection of attribute extents and every concept intent
is the intersection of object intents.
A ∩ m0
and include it into the list, provided that it is not yet contained within it.
It is easy to see that this list finally contains exactly those sets which can be writ-
ten as an intersection of attribute extents, i.e. according to the proposition exactly
the concept extents. Then, using the formal context, we can find the concept in-
tent A0 for each such extent A. Thus we obtain a list of all formal concepts (A, A0 )
of the context.
Figure 2.1 List of concept extents for the formal context in Figure 1.1.
Determining the concepts in this way is usually easier if one starts drawing a
diagram of the concept lattice at the same time. We show this, staying with the
example of the “living beings and water” context from Figure 1.1. The interme-
diate steps are presented in Figure 2.2. First, draw a small circle for the largest
concept in the context. If there are attributes whose attribute extent contains all of
G, then these attribute names are entered above that circle. In our example, this
would be “a”. Then we determine the attributes the extents of which are maxi-
mal among the remaining attribute extents. In our example, we obtain b, c, d and
g. The attribute concept of each of these attributes is represented by a small circle
below the circle already drawn. These circles are then linked to the circle of the
largest concept and the names of the new attributes are entered above the respec-
tive circles. Now we systematically form the infima of the already represented
2.1 All formal concepts of a formal context 79
ba −→ ba −→ ba
HH H
J H J HH
g g
b b JJ bb HH bd
c
b cb JJ bb HH bd
J
JJ b
3,6
−→ ba −→ ba
HH HH
J H J H
g g
b c b JJ bb HH bd b c b JJ bb HH bd
H H HH HH
J H HH J
H J J J
JJ b HH b HH JJ b JJ JJ HHH HHH
6,7,8
b3,4
b b
1,2,3 H 6,7,8
b JJ b
5,6 5,6
3,6 H 3,6 H J
H
HH J J HH
HJJ b JJ b JJ b
H
6 3 H 6
HH
HH b
−→ ba −→ ba
J H
HH H
J HH
g g
b b JJ bb HH bd b b JJ bb HH bd
c c
hb J
H HH J bf hb J
JH
H H
JH
H HH J bf
H H
b J JJ b JJ b HH b HH Jb b
J JJ b JJ b HH b HH Jb
3,4 1,2,3 H 6,7,8 5,6
J J b2,3 3,6
H J b 6,8 J Jb H Jb
J H J HH
iqb JJ b eb HHJb ibq JJ b eb HHJb
4 3 H 7 6 H
HH HH
HH qb HH qb
Figure 2.2 Intermediate steps in the drawing of Figure 1.3 (p. 29). The last step is to insert
the object names, which also helps verifying whether the diagram is correct.
concepts and represent the newly generated concepts by small circles with their
respective connecting lines. In our example, this procedure is first applied to the
concepts for b and c, then to those for b, c and d, and finally to those for b, c, d
and g, making use of our knowledge about concepts already determined (if nec-
essary one can note down the extent intersections temporarily at the respective
circles in order to remember them for later). If we have drawn the line diagram
for all concepts determined by this stage, we look for the attributes the extents of
which are maximal among the attribute extents not already used. In our exam-
ple, we obtain e, f and h. As above, we represent the attribute concepts and all
new intersections of the concepts now available. In our example, we would have
to go through this procedure one more time, that is for the attribute i. If we fi-
nally have worked our way through all the attributes, the resulting line diagram
80 2 Determination and representation
The algorithm described above for determining all concepts becomes cum-
bersome for larger contexts, since the created list must be consulted again and
again. There are faster algorithms for the generation of all concept extents, and
we have already given one in Section 0.3: “Next closure” (Algorithm 1, p. 12).
In fact, as a direct consequence of Proposition 12, it had already been stated that
the concept extents form a closure system and that the associated closure opera-
tor is the “extent closure”, i.e., the mapping X 7→ X 00 . Therefore, the algorithm
can be applied, and Algorithm 3 below is identical to Algorithm 1 except for an
adjustment to the symbols. (In the same way, of course, it applies to the closure
system of the concept intents (Next intent), since the intents of K are exactly
the extents of Kd . But it is not necessary to do both, because the intents can di-
rectly be computed from the extents.) To compute all concept extents, start with
the “first extent” A := ∅00 and then invoke Algorithm 3 until ⊥ is returned. The
previous output is used as input A in each such step.
The algorithm assumes a linear order of the set of objects G. However, this is
arbitrary and can be freely chosen, which occasionally results in additional op-
timization possibilities. For simplicity, we assume G = {1 < 2 < . . . < n}.
We consider the set of all subsets of G to be “in lexicographic order”, which
is defined next. A subset A ⊆ G is called lectically smaller than a subset B 6= A,
if the smallest element distinguishing A and B from each other belongs to B.
Formally:
This defines a linear strict order for the powerset P(G), i.e., for subsets A 6= B
always A < B or B < A holds. The algorithm finds for an arbitrary given set
A ⊆ G the extent that is smallest after A with respect to this lectic order. The
lectically smallest concept extent is ∅00 . The other extents are found incrementally
by determining the one which is lectically next to the last extent found. In the end,
we obtain the lectically largest extent, namely G. Figure 2.3 shows the result in
the case of the “living beings and water” context.
To make this precise, we define for A, B ⊆ G, i ∈ G,
A <i B :⇔ i ∈ B \ A and A ∩ {1, 2, . . . , i − 1} = B ∩ {1, 2, . . . , i − 1}.
A ⊕ i := ((A ∩ {1, 2, . . . , i − 1}) ∪ {i})00 .
It is easy to verify the following statements:
(1) A < B ⇔ A <i B for one i ∈ G.
(2) A <i B and A <j C with i < j ⇒ C <i B.
(3) i 6∈ A ⇒ A < A ⊕ i.
(4) A <i B and B extent ⇒ A ⊕ i ⊆ B, thus A ⊕ i ≤ B.
(5) A <i B and B extent ⇒ A <i A ⊕ i.
Theorem 5 The smallest concept extent larger than a given set A ⊂ G (with respect to
the lectic order) is
A ⊕ i,
i being the largest element of G with A <i A ⊕ i.
Proof Let A+ be the smallest extent after A with respect to the lectic order. On
account of A < A+ , we get A <i A+ for some i ∈ G by (1) and thus A <i A⊕i by
(5). By (4) it follows that A ⊕ i ≤ A+ , i.e., A ⊕ i = A+ because of A < A ⊕ i. The
fact that i is the largest element with A <i A⊕i results from (2), since A <j A⊕j
with j 6= i on account of A ⊕ i = A+ < A ⊕ j by (2) yields j < i.
Figure 2.3 List of the extents for the context in Figure 1.1 in lectic order. Behind each
extent A, the element i with A+ = A ⊕ i is stated.
It is not difficult to modify the Next extent algorithm so that it also generates
the neighborhood relation of the concept lattice, which is needed for the concept
lattice diagram. However, this requires that a list of concepts is created. See [154],
Algorithms 11 and 12.
Because the algorithm is so simple, it allows numerous variations and exten-
sions. We briefly mention an example and then formulate a theorem that extends
Theorem 5. Many other results can be found in [154].
We can take advantage of the fact that the extents are generated in lectic order.
For example, if
C := {1, 2, . . . , c}, D := {c + 1, c + 2, . . . , d} ⊆ G,
then we obtain as the lectic successors of C first all sets which contain C and are
disjoint to D. For arbitrary disjoint subsets C, D ⊆ G, one can find all concept
extents E of (G, M, I) with C ⊆ E, E ∩ D = ∅ by suitably adjusting the linear
order of G.
The premises of the algorithm can be weakened in some respects, which al-
lows generalizations. Without substantial modifications of the proof we obtain
the following theorem.
Theorem 6 If F is a family of extents of the context (G, M, I) with the property
we obtain for an arbitrary subset A ⊆ G the set A+ which is the lectically next in F –if
it exists– by
A+ = A ⊕ i,
The best and most versatile form of representation for a concept lattice is a well
drawn line diagram. It is however tedious to draw such a diagram by hand and
one would wish an automatic generation by means of a computer. Both the avail-
able algorithms and the implementations are gradually getting better and better,
but there is still no program that provides a general satisfactory solution. It is by
no means clear which qualities make up a good diagram. It should be transparent,
easily readable and should facilitate the interpretation of the data represented.
How this can be achieved in each individual case depends however on the aim of
the interpretation and on the structure of the lattice. Simple optimization crite-
ria (minimization of the number of edge crossings, drawing in layers, etc.) often
lead to unsatisfactory results. Nevertheless, automatically generated diagrams
are a great help: they can serve as the starting point for drawing by hand. There-
fore, we will describe simple methods of generating and manipulating line dia-
grams by means of a computer, later we suggest even better procedures with the
aid of the structure theory for concept lattices.
b: not equilateral
a: equilateral
c: isosceles
d: oblique
f: obtuse
e: acute
g: right
As an illustration, we will use the formal context in Figure 2.4, in which trian-
gles are classified according to properties such as right-angled, equilateral, etc. The
choice of the triangles is not coincidental: the context is the result of an attribute
exploration, a technique to be discussed in the next section. But for the moment
we are only concerned with the question of how to obtain a line diagram for this
context.
We can use a computer program to obtain the concepts of the formal context
and the edges of the line diagram. The successor list is displayed on the right. We
84 2 Determination and representation
can read from it that the context has 18 concepts. These are denoted by the serial
numbers 1,. . . ,18. Behind the colon follow the upper neighbors of each concept.
In the line diagram, an edge must be drawn to each of the upper neighbors, and
those are all edges.
Obviously concept no. 1 is the unit element of
the concept lattice (since it has no upper neigh-
bor) and no. 18 is the zero element (since 18 does
not occur as an upper neighbor). 1: −
2: 1 −
As a graph, the line diagram is already com- 3: 2 −
pletely determined by this list. It can be used to 4: 1 −
sketch a diagram “from bottom to top”: first of all 5: 2 4 −
we draw the smallest element (concept no. 18), 6: 3 5 −
above it the upper neighbors (13, 15, 16, 17), then 7: 1 −
8: 7 −
their upper neighbors (6, 8, 10, 11, 12, 14), and so 9: 2 7 −
on. It is still open how points are arranged on pa- 10 : 9 −
per. This can be done “intuitively” but will then 11 : 3 9 −
require various iterations to develop a satisfactory 12 : 4 7 −
diagram. 13 : 8 12 −
14 : 5 9 12 −
There is however an efficacious method to sup- 15 : 10 14 −
port the generation of a line diagram. This geo- 16 : 6 11 14 −
metrical method is based on first understanding 17 : 6 −
the lattice-theoretic structure through a geometri- 18 : 13 15 16 17 −
cal representation of the concept lattice and then
to find the best possible arrangement for the line
diagram. This means that we draw as an interme-
diate step –by hand or with the aid of a computer– an auxiliary picture, which is
then used to draw the actual line diagram. This auxiliary picture is called the ge-
ometrical diagram. Intuitively, we think of this diagram in the following way: we
imagine that the lattice is realized by means of a three-dimensional line diagram
and look down on the lattice from its highest point, i.e., from the unit element.
From the top, we first see the lower neighbors of the unit element. In the ge-
ometrical diagram they are represented by unconcealed circles into which we
write the names of the respective elements. We continue to draw the geometrical
diagram in accordance with the following rules:
1. An element with exactly one upper neighbor is represented by a circle which
is partly covered by the upper neighbor.
2. An element with exactly two upper neighbors is represented by a connecting
line segment between the two upper neighbors. The name of the element is
written into a circle which is partly covered by this connecting line.
3. An element with exactly three upper neighbors is represented by a connecting
triangle between the upper neighbors. The name of the element is written into
the triangle.
2.2 Diagrams of concept lattices 85
In this way we obtain the geometrical diagram in Figure 2.5. The individual
steps are noted down in the following table. The necessary information has been
taken from the above successor list.
2 lies immediately below 1: therefore a circle for 2.
3 immediately below 2: therefore a circle for 3, partly covered by the 2-circle.
4 immediately below 1: therefore a circle for 4.
5 immediately below 2 and 4: therefore a line segment for 5 between the 2-circle and the
4-circle.
6 immediately below 3 and 5: therefore a line segment for 6 between the 3-circle and the
5-line-segment.
7 immediately below 1: therefore a circle for 7.
8 immediately below 7: therefore a circle for 8, partly covered by the 7-circle.
9 immediately below 2 and 7: therefore a line segment for 9 between the 2-circle and the
7-circle.
10 immediately below 9: therefore a circle for 10, partly covered by the 9-circle.
11 immediately below 3 and 9: therefore a line segment for 11 between the 3-circle and the
9-line-segment.
12 immediately below 4 and 7: therefore a line segment for 12 between the 4-circle and the
7-circle.
13 immediately below 8 and 12: therefore a line segment for 13 between the 8-circle and the
12-line-segment.
14 immediately below 5, 9 and 12: therefore a triangle for 14 between the 5-circle, the 9-
circle and the 12-line-segment.
86 2 Determination and representation
15 immediately below 10 and 14: therefore a line segment for 15 between the 10-circle and
the 14-triangle.
16 immediately below 6, 11 and 14: therefore a triangle for 16 between the 6-line-segment,
the 11-line-segment and the 14-triangle.
17 immediately below 6: therefore a circle for 17, partly covered by the 6-line-segment.
d1
2d d4@ d7
@
d 5@ d9@ d12@
@ @ @
3d
@ @ @ @ @
6 d @ d11@ d14 @d @ @d
10 8
@ @ @
17 d
@d @d @d
16 15 13
@ B
@ B
@
B
@ B
@B
@B d
18
It still remains to be said how a good line diagram can be obtained from the
geometrical diagram. The derived line diagram for the concept lattice of the tri-
angles is presented in Figure 2.6. If one already has some experience with the
geometrical method, one can see from Figure 2.5 that the most striking substruc-
ture of the lattice consists of two Boolean cubes. But even without this experience,
one can soon reach this conclusion by proceeding systematically. As a rule, one
should start with the lower neighbors of the unit element being represented by
unconcealed circles. In Figure 2.5 these are the 2-, 4- and 7-circle. These circles are
connected pairwise by the line segments 5, 9 and 12, which in turn are connected
by the 14-triangle. This shows that the concepts 1, 2, 4, 5, 7, 9, 12 and 14 form a
Boolean sublattice. The question is how these eight elements can best be arranged
within the line diagram. After having drawn the unit element, it seems advisable
to put the co-atoms 2 and 7 on the outside and 4 between them, since below both
2 and 7 there is another “point”, which needs some space. The concepts 5, 7, 9
and 12 will be best placed in accordance with the rule of parallelograms, which says
that one should (if possible) place an element in a way that it makes up a paral-
lelogram together with three elements already represented and their connecting
line segments. The resulting picture of the Boolean sublattice represents a cube
standing on one of its corners. After the explanation given so far, it should not
be difficult to recognize the second Boolean sublattice, consisting of the concepts
2, 3, 5, 6, 9, 11, 14 and 16. Since the cube representing it shares the elements 2,
2.2 Diagrams of concept lattices 87
5, 9 and 14 with the first cube, it is obvious how to continue the drawing. How-
ever, one further rule should be observed, the so-called rule of lines, according to
which a line to a new “point” should be arranged in such a way that it continues
some line segments already drawn. If we observe the rule of lines and the rule of
parallelograms for the remaining elements 8, 10, 13, 15 and 17, we obtain from
the geometrical diagram a satisfactory line diagram, to which we only have to
add the zero element (no. 18) (cf. Figure 2.6). For the labelling with object and
attribute names, additional information is required, which a computer program
could supply by means of an assignment list (see Fig. 2.7).
d isosceles
oblique
d d%@ dnot equilateral
@ %
acute
d d @ d @ d @ obtuse
@ @ @
d @ d @ d @ d @ @ dright
@ @ @ @ @
equilateral @ @ @
d @d @d @d
@ B
@ B
@
B
@ B
@B
@B d
diagrams can still be used as instructions for drawing good line diagrams. It is
helpful to observe geometrical patterns and their respective realizations in the
line diagrams. In some (relatively rare) cases, it is advisable to construct the line
diagram from bottom to top; in this case one should use the so-called predecessor
list.
Both of the procedures described above make use of the computer in order
to obtain information necessary for a diagram. We will now explain a method
where a computer generates a diagram and offers the possibility of improving it
interactively. Programming details are irrelevant in this context. We will there-
fore only give a positioning rule which assigns points in the plane to the ele-
ments of a given ordered set (P, ≤). If a and b are elements of P with a < b,
the point assigned to a must be lower than the point assigned to b (i.e., it must
have a smaller y-coordinate). This is guaranteed by our method. We will leave
the computation of the edges and the checking for undesired coincidences to the
programming. We do not even guarantee that our positioning is injective (which
of course is necessary for a correct line diagram). This must also be checked if
necessary.
Definition 39 A set representation of an ordered set (P, ≤) is an order embed-
ding of (P, ≤) in the powerset of a set X, i.e., a map
rep : P → P(X)
X := G, (A, B) 7→ A
resp. X := M, (A, B) 7→ M \ B
are representations which can be combined to
X := G ∪˙ M, (A, B) 7→ A ∪ (M \ B).
vec : X → R2 ,
large small
natural artificial
flowing stagnant
Figure 2.9 An additive line diagram of the concept lattice of a lexical field “waters”. The set
representation is based on the irreducible attributes, i.e. the positioning of the attribute
concepts determines that of all remaining concepts. If we interpret the line segments be-
tween the unit element and the attribute concepts as vectors, we obtain the position of an
arbitrary concept by the sum of the vectors belonging to attributes of its concept intent
starting from the unit element. Other diagrams for the same lattice can be found in Figure
2.10.
Every finite line diagram can be interpreted as an additive diagram with re-
spect to an appropriate set representation. For concept lattices we usually use
the representation by means of the irreducible objects and/or attributes. The re-
sulting diagrams are characterized by a great number of parallel edges, which
90 2 Determination and representation
Figure 2.10 Line diagram and nested line diagram. Less trivial examples of nested line
diagrams can be found in Figures 1.15 on p. 50 and 2.17 on p. 103.
Proof If (A, B) is a concept of (G, M, I), then B ∩ Mi is the set of all attributes
common to the objects of A in the context (G, Mi , I ∩ G × Mi ), i.e., it is an intent
of this context. Hence the above-mentioned assignment is really a map into the
product. The union of the intents B ∩ M1 and B ∩ M2 again yields B, i.e., the
92 2 Determination and representation
map is injective. The fact that it is furthermore -preserving (and thus an order-
W
embedding) can again be seen from the concept intents. It remains to be shown
that the component maps are surjective. Let C be an intent of (G, Mi , I ∩G×Mi ).
Then B := C II is an intent of (G, M, I) with B ∩ Mi = C, i.e., the image of the
concept (B 0 , B) of (G, M, I) under the ith component map is the concept with
the intent C.
In order to sketch a nested line diagram, we proceed as follows: First of all we
split up the attribute set: M = M1 ∪ M2 . This splitting up does not have to be
disjoint. More important for interpretation purposes is the idea that the sets Mi
bear meaning. Now, we draw line diagrams of the subcontexts Ki := (G, Mi , I ∩
G × Mi ), i ∈ {1, 2} and label them with the names of the objects and attributes,
as usual. Then we sketch a nested diagram of the product of the concept lattices
B(Ki ) as an auxiliary structure. For this purpose we draw a large copy of the
diagram of B(K1 ), representing the lattice elements not by small circles but by
congruent rectangular boxes, each of which contains a diagram of B(K2 ).
WBy Theorem 7 the concept lattice B(G, M, I) is embedded in this product as
a -semilattice. If a list of the elements of B(G, M, I) is available, we can enter
them into the product according to their intents. If not, we enter the object con-
cepts the intents of which can be read off directly from the context, and form all
suprema.
This at the same time provides us with a further, quite practicable method of
determining a concept lattice by hand: split up the attribute set as appropriate,
determine the (small) concept lattices of the subcontexts, draw their product
in form of a nested line diagram, enter the object concepts and close it against
suprema. This method is particularly advisable in order to arrive at a useful di-
agram quickly.
2.3 Implications
described in the previous section. In such cases, the conceptual structure can be
inferred from the implications between the attributes, by which we mean statements
of the following kind:
“Every object with the attributes a, b, c, . . . also has the attributes x, y, z, . . .”
Formally, an implication between attributes (over M ) is a pair of subsets of the
attribute set M . It is denoted2 by A → B. The set A is the premise and B is the
conclusion.
In this section, we examine the set of all attribute implications that hold in
a given formal context. It turns out that these implications contain enough in-
formation to reconstruct the structure of the concept lattice. Conversely, the at-
tribute implications of a formal context can also be read from the concept lattice.
However, the system of all implications between attributes holding in a given
context tends to be large and to contain many trivial implications. Therefore, we
look for subsystems which suffice to describe the concept lattice. We start with
some simple definitions:
Definition 40 A subset T ⊆ M respects an implication A → B if A 6⊆ T or
B ⊆ T . T respects a set L of implications if T respects every single implication
in L. A → B holds in a set {T1 , T2 , . . .} of subsets if each of the subsets Ti respects
the implication A → B. A → B holds in a formal context (G, M, I) if it holds in
the system of object intents. In this case, we also say that A → B is an implication
of the context (G, M, I) or, equivalently, that within the context (G, M, I), A is a
premise of B. ♦
Proposition 22 An implication A → B holds in (G, M, I) if and only if B ⊆ A00 . It
then automatically holds in the set of all concept intents as well.
How can we read off an implication from a concept lattice diagram? It is suf-
ficient to describe this for implications of the form A → m, since A → B holds
if and only if A → m holds for eachVm ∈ B. Now A → m holds if and only
if (m0 , m00 ) ≥ (A0 , A00 ), i.e., if µm ≥ {µn | n ∈ A}. This means that we have
to check in the diagram whether the concept labeled by m is located above the
infimum of all concepts labeled by an n from A.
It can occasionally be useful for the determination of the implications to re-
place the original context (G, M, I) by its complementary context (G, M, (G ×
M ) \ I), in particular if the latter has considerably fewer concepts as (G, M, I).
For m ∈ M and T A 0⊆ M the following S hold:
equivalences m ∈ A00 ⇔ {m} ⊆ A00
⇔ A ⊆ m ⇔ {n | n ∈ A} ⊆ m ⇔ G \ m ⊆ {G \ n | n ∈ A}. Thus, A → m
0 0 0 0 0
holds in the context (G, M, I) if and only if in the complementary context every
object with the attribute m has at least one attribute n from A.
Proposition 23 If L is a set of implications over M ,
Mod(L) := {X ⊆ M | X respects L}
2 When the sets are small, we shall omit brackets (as we have done earlier), i.e., we shall write
A → m instead of A → {m}, etc.
94 2 Determination and representation
We form the sets X L , X LL , X LLL , . . . until we finally obtain a set L(X) := X L...L
with L(X)L = L(X) (in the case of infinite contexts it can be necessary to con-
tinue this process transfinitely). L(X) then is the closure of X with respect to the
closure system Mod(L) which we have been looking for.
By means of the closure system Mod(L) it is also possible to construct a con-
text for every given set L of implications, the intents of which are precisely the
sets respecting L: (Mod(L), M, 3) has this property. In addition to L in this con-
text all implications hold which follow from L in the sense of the following def-
inition:
Definition 41 An implication A → B follows (semantically) from a set L of
implications between attributes if each subset of M respecting L also respects
A → B. A family of implications L is called closed if every implication following
from L is already contained in L.
A set L of implications of a formal context (G, M, I) is called complete if every
implication of (G, M, I) follows from L.3 ♦
In other words: An implication follows semantically from L if it holds in every
system of sets in which L holds as well. This is the case if and only if Mod(L) =
Mod(L ∪ {A → B}). More about this in the following section.
There is also a syntactic characterization of closed implication sets known as
the Armstrong rules [5]:
B ⊆ L(A)
holds, using a fast algorithm for computing L(A), e.g., Algorithm 15 in [154].
3 Note that a complete set must always consist of implications which hold in (G, M, I).
2.3 Implications 95
It should be noted that sometimes the inference rules can change, namely
when additional “background knowledge” about the attributes is known. A typ-
ical instance is that one attribute is known to be the “negation” of another. This
will be discussed in Section 8.1.
the set of those attributes contained in A00 but not in A or in the closure of any
proper subset of A. We call A a proper premise if A• 6= ∅, i.e., if
[
A00 6= A ∪ (A \ {n})00 .
n∈A
In certain respects, the set of proper premises is canonical with respect to the
property of being iteration-free as described in Proposition 25. In order to state
this more precisely, we first introduce a further term. A family of implications
can be simplified by merging implications with the same premise. We call a fam-
ily of implications contracted if there are no premises which occur more than
once. If L is any contracted family of implications satisfying the condition of the
proposition, i.e., with
[
T 00 = T ∪ {B | A → B ∈ L, A ⊆ T } for all T ⊆ M,
(M \ g 0 ) ∩ P 6= ∅
holds for all g ∈ G with g . m. P is a proper premise for m if and only if m 6∈ P and P
is minimal with respect to the property that (M \ g 0 ) ∩ P 6= ∅ holds for all g ∈ G with
g . m.
Proof For g ∈ G and P ⊆ M we have the equivalences
(M \ g 0 ) ∩ P 6= ∅ ⇔ P 6⊆ g 0 ⇔ g 6∈ P 0 .
plications. For the following results, we make the general assumption that the
attribute set M which occurs is finite. This permits a recursive definition of the
basic notion of a pseudo-intent (which takes the place of the proper premise):
Definition 44 P ⊆ M is called a pseudo-intent of (G, M, I) if and only if P 6= P 00
and Q00 ⊆ P holds for every pseudo-intent Q ⊆ P , Q 6= P . ♦
L := {P → P 00 | P pseudo-intent}
The canonical basis is indeed canonical with respect to the properties stated.
For this we start with a simple proposition:
Proposition 27 If P and Q are intents or pseudo-intents with P 6⊆ Q and Q 6⊆ P , then
P ∩ Q is an intent.
Proof P as well as Q and thus also P ∩ Q respect all implications in L with the
possible exception of P → P 00 and Q → Q00 . If P 6= P ∩ Q 6= Q, then P ∩ Q also
respects these implications, i.e., it is an intent.
4 Other names are “Duquenne-Guigues basis” or “stem base”.
98 2 Determination and representation
The next proposition shows among other things that there can be no complete
set which contains fewer implications than there are pseudo-intents:
Proposition 28 Every complete set Σ of implications contains, for each pseudo-intent
P , an implication A → B with A ⊆ P and A00 = P 00 .
Proof A pseudo-intent P is always not equal P 00 . Therefore, provided that Σ is
complete, there must be at least one implication A → B in Σ which leads out of
P , i.e., with A ⊆ P and B 6⊆ P . On account of B ⊆ A00 , we get A00 6⊆ P , and thus
A00 ∩ P cannot be a concept intent. By Proposition 27 this yields P ⊆ A00 and thus
P 00 = A00 .
The recursive definition of the pseudo-intents provides us with a first, al-
though inefficient, algorithm for generating them. In the following we will de-
velop a more practicable procedure. As an immediate consequence of Proposi-
tion 27 we obtain:
Attribute exploration
We now return to the initial question of this section: how can we determine the
concept intents using the implications? We have seen that we do not need all
implications for this, but that a small subset of them is sufficient. So far, we have
2.3 Implications 99
A := ∅
L := ∅
while A 6= M do
if A 6= A00 then
L := L ∪ {A → A00 }
A := Next Closure(A, M, L)
return L
only explained how to obtain these implications from an existing formal context.
However, using the tools now at our disposal, we can also develop a method to
develop implication lists that are free of redundancy, even if the formal context
is only partially available.
This procedure, called attribute exploration, has been proven in many appli-
cations. It is documented in a separate monograph [154]. In practice, a computer
is used to manage the implication sets and to calculate which information is still
missing. The implications are interactively determined, i.e. in cooperation with
the user.
The algorithm for determining the pseudo-intents permits a modification re-
sulting in an interactive program: it is possible to modify the formal context by
adding new objects, even while the generation of the list L of the implications is
in progress. If the intents of these objects respect all implications determined so
far, the computation for the new context can be continued with the results so far
obtained. This is the content of the following proposition:
Proposition 30 Let K be a formal context and let P1 , P2 , . . . , Pn be the first n pseudo-
intents of K with respect to the lectic order. If K is extended by an object g the object
intent g 0 of which respects the implications Pi → Pi00 , i ∈ {1, . . . , n}, then P1 , P2 , . . . ,
Pn are also the lectically first n pseudo-intents of the extended context.
This can be proved for example by induction on n.
Thus, if we have found a new pseudo-intent P , we can stop the algorithm
and ask, whether the implication P → P 00 should be added to L. The user can
answer this question in the affirmative or add a counter-example, which must not
contradict the implications which were confirmed so far. In the extreme case, the
procedure can be started with a context the object set of which is empty. In this
case, the user will have to enter all counter-examples, thereby creating a concept
system with a given “attribute logic”.
A detailed description of the procedure can be found in [154]. We limit our-
selves here to an example. From a book on measurement theory [321] we take
a list of properties of binary relations, which are used there in order to define
different types of relations, see Figure 2.12.
100 2 Determination and representation
Property Definition
r reflexive xRx for all x ∈ S
i irreflexive ¬xRx for all x ∈ S
s symmetric xRy ⇒ yRx for all x, y ∈ S
as asymmetric xRy ⇒ ¬yRx for all x, y ∈ S
an antisymmetric xRy and yRx ⇒ x = y for all x, y ∈ S
t transitive xRy and yRz ⇒ xRz for all x, y, z ∈ S
nt negatively transitive ¬xRy and ¬yRz ⇒ ¬xRz for all x, y, z ∈ S
c connex xRy or yRx for all x 6= y ∈ S
sc strictly connex xRy or yRx for all x, y ∈ S
Which implications exist between those properties? For every such implica-
tion it is easy to determine whether it holds for all binary relations, and there can
only be finitely many (since we are dealing with a finite number of attributes),
but very likely there are more than we would like to list. Our algorithm should
allow us to find only “good” implications right away. Implications which are not
valid for all binary relations have to be refuted by stating counterexamples.
First, we equip ourselves with a small sample set, for which we take all re-
lations on a one- or two-element set. Up to isomorphism there are twelve such
relations (Figure 2.13).
S R r i s as an t nt c sc
{0} ∅ × × × × × × ×
{0} {(0, 0)} × × × × × × ×
{0, 1} ∅ × × × × × ×
{0, 1} {(0, 0)} × × × ←
{0, 1} {(0, 0), (1, 1)} × × × ×
{0, 1} {(0, 0), (0, 1)} × × × × ←
{0, 1} {(0, 0), (1, 0)} × × × × ←
{0, 1} S × S \ {(0, 1)} × × × × × ×
{0, 1} S × S \ {(0, 0)} × × × ←
{0, 1} {(0, 1)} × × × × × ×
{0, 1} {(0, 1), (1, 0)} ×× × ×
{0, 1} S×S × × × × × ×
Now we have a context to start with (generally, this context can even be
empty). Of course, only implications which hold in this context can hold for all
binary relations, but not vice versa. Please note that the four objects marked by
2.3 Implications 101
No. S R
1 {0} ∅
2 {0} {(0, 0)}
3 {0, 1} ∅
4 {0, 1} {(0, 0), (1, 1)}
5 {0, 1} S × S \ {(0, 1)}
6 {0, 1} {(0, 1)}
7 {0, 1} {(0, 1), (1, 0)}
8 {0, 1} S×S
9 {0, 1, 2} S × S \ {(0, 1), (1, 2), (2, 0)}
10 {0, 1, 2} {(0, 1), (1, 2), (2, 0)}
11 {0, 1, 2} {(0, 1)}
12 {0, 1, 2} {(0, 1), (1, 0)}
13 {0, 1, 2} S × S \ {(0, 1)}
14 {0, 1, 2} S × S \ {(0, 1), (1, 0)}
a ← are superfluous, since their intents are the intersections of other object in-
tents and therefore respect all implications respected by the other objects. We
will leave them out in the following. Now we use the algorithm in order to cal-
culate the first pseudo-intent. The lectically smallest pseudo-intent in this context
is {sc}, with {sc}00 = {r, t, nt, c, sc}. In other words, the implication
holds in all examples stated so far. Does it hold for binary relations in gen-
eral? Of course not. A counter-example is for instance S = {0, 1, 2}, R =
S × S \ {(0, 1), (1, 2), (2, 0)}. This relation is reflexive, antisymmetric, connex
and strongly connex and has none of the other properties. We add this exam-
ple to our formal context and again ask for the smallest pseudo-intent. It is still
{sc}, but now {sc}00 = {r, c, sc}, and we have to check whether the implication
{sc} → {r, c, sc}, which we abbreviate by
sc → r, c,
holds for all binary relations. As a matter of fact, every strongly connex relation
is reflexive and connex. Therefore, we can add this implication to the list of im-
plications L.
The next pseudo-intent is {t, c} with {t, c}00 = {t, nt, c}. This suggests the im-
plication
t, c → nt,
which in fact holds for binary relations and therefore is added to the list, as well
as the following one
102 2 Determination and representation
r i s as an t nt c sc
1 × × × × × × ×
2 × × × × × × ×
3 × × × × × ×
4 × × × ×
5 × × × × × ×
6 × × × × × ×
7 × × × ×
8 × × × × × ×
9 × × × ×
10 × × × ×
11 × × × ×
12 × ×
13 × × × ×
14 × ×
Figure 2.16 A complete and non-redundant list of implications, given by the canonical
basis of the formal context in Figure 2.15, slightly simplified for better readability.
2.3 Implications 103
symmetric antisymmetric
A e.Q
A
reflexive ` Q
` eA e AA eQ e irreflexive
Q
Q Q
A A Q A
eA e A Q eQA e
14 A
Q Z
Q AZ
A
Q A 12
Q
Q
Q
@
transitive negatively transitive @
connex
@
e e e
Q
AQ
Q Q
AQ
e e A e Q e A Q e e A eQ e
AQ
Q QQ Q Q
A AQ QA AAQ
Q A A Q A
eA eA e Q QA e A A Q e Q A e A Q QA e
Q A Q
A
Q A 11
Q A Q
QA A 9 QQ A 10
e
AA
QA A
Q
A A
Q
A
4 QQ
Q Q
Q Q
Q Q Q
@ @
@ @
@ @
e e
AQ AQ
Q
e A e Q e
e A Q e
Q
Q 13 Q Q
A AQ
QA AQ Q
A A e Q QA e A A Q e7Q
Q
A A
A
A Q QA A Q
QA e QA
A
A
Q
A
Q 3
Q Q
Q
Q Q
@
@
@
e
Q
e e A e Q
AQ
Q Q
A A QA
e5A eA e Q QA e
Q
8 AA Q QA 6
e QA e
2 QQ 1
Q e
Figure 2.17 The concept lattice for the context of the binary relations. The numbers refer to the
list of relations in Figure 2.14.
104 2 Determination and representation
an, nt → t,
which results from the pseudo-intent {an, nt} with
as → i, an, t, nt
does not hold generally, as the following example shows: S := {0, 1, 2}, R :=
{(0, 1), (1, 2), (2, 0)}. This relation has the attributes i, as, an, nt, and we add it
to the context. Since it obviously respects all implications accepted so far, it has
no consequences for the pseudo-intents found up to then (cf. Proposition 30).
In the following, we first confirm the implications as → i, an and s, c → nt
as well as s, an → t, then we state a counter-example for i, t → as, an, nt etc.
The complete result is presented in Figures 2.14 to 2.17. We point out that the
premises of the implications in Figure 2.16 are precisely the pseudo-intents of
the context in Figure 2.15.
The procedure does not guarantee that the resulting formal context is reduced
(as in the example). Newly entered objects can make previously entered objects
dispensable. It is possible to “row-reduce” the context during the process (i.e.,
to delete dispensable objects). This has no effect on the implications.
How can the theory of implications be applied to the case of many-valued con-
texts? There are several answers to this question.
One is to study one-valued contexts that are derived from a many-valued con-
text through conceptual scaling, as introduced in Section 1.4. Basically, that leads
to implications between the individual attribute values, at least as long as we
keep to plain scaling. A typical such implication might say that
“Downtown real estate is expensive”.
However, the theory needs modification because the conceptual scales may pro-
vide additional background knowledge, resulting in a modified, stronger impli-
cation inference. This will be discussed in the second subsection, after we have
compiled some facts about clauses, which are a useful tool.
In colloquial language the term dependency of many-valued attributes is used
as exemplified by the following sentence
2.4 Clauses and dependencies 105
Clauses
{needs water to live, has limbs, can move around, lives on land}.
The fact that no object label is attached to this concept means that no object in
the formal context has these four attributes, but no others. This concept intent
consists of the common attributes of dog and frog, but is not represented by a
single object from G. And every single attribute that dog and frog have in com-
mon, they share with another living being from G. Only the attribute combination
is characteristic for dog and frog. If, for example, “lizard” were added as a new
(reducible) object, then the dog-frog-lizard concept would get an object label as
well.
According to Proposition 23 (p. 93), the implications of a formal context K de-
termine the closure system of its concept intents and thus the concept lattice up
to isomorphism. However, they do not determine the formal context, not even up
to isomorphism, since adding reducible objects to K does not change its impli-
cations. The implications do not provide any information about which concept
intents are also object intents. This is only clear for -irreducible concept intents,
W
which must be object intents.
A formal concept without an object label leads to an expression similar to an
implication:
Every object in this concept extent has an attribute
which is not in the intent.
In the example discussed above, this specializes to
Every living being from G that lives on land and can move around
suckles its offspring or also lives in water.
The decisive
V difference to an implication is the word “or”. Such expressions of
the form A → S, called clauses over M , are dealt with in Section 8.1.
W
Clauses are much more expressive than implications. Theorem 67 (p. 314)
will show that every formal context is determined by its clauses up to clarifica-
tion and isomorphism. But clauses are also much more difficult to handle than
implications, both from an algorithmic point of view and for human intuition.
Nevertheless, clauses can be very useful. Among other things, they provide ad-
ditional options for dealing with implications. We start with a special case.
Suppose you know that two attributes m and n are negations of each other.
On the one hand, this means that no object can have both attributes, but it also
means that every object must have one of them. If the implications A ∪ {m} → s
and A ∪ {n} → s both hold for some A ⊆ M and s ∈ M , then A → s must
also hold, obviously. But that can not be concluded using only the Armstrong
rules (see page 94). If two attributes are negations of each other, this results in
additional inference rules for implications. The same works for every clause.
Proposition 31 The following are equivalent:
1. A → S is a clause of K.
V W
2. The set L of implications of K is closed under the inference rule
If A ⊆ X and X ∪ {s} → B ∈ L for all s ∈ S, then
R∧A→∨S
X → B ∈ L.
Proof The inference rule R∧A→∨S does not hold for the implications of K iff
there is some attribute m ∈ M and some subset X ⊆ M containing A such that
all implications X ∪ {s} → m are in L, but X → m is not. X → m is not in L iff
there is some object g of K such
V that W
g 0 does not respect X → m, i.e., with X ⊆ g 0
and m ∈ / g . But g respects A → S if and only if A ⊆ g 0 enforces s ∈ g 0 for
0 0
some s ∈ S.
Remark: The inference rules R∧A→∨S are introduced here as a tool of theory.
For the algorithmic implementation of implication inference with background
knowledge, we refer to Section 5.3 of [154].
Theorem 9 Let K := (G, M, I) be a formal context with finite attribute set M . A set L
of implications over M is the set of implications of some subcontext (H, M, I ∩(H ×M ))
if
Vand onlyW if L is closed under the Armstrong rules and under all rules R∧A→∨S , where
A → S is a clause of K.
{g 0 \ A | A ⊆ g 0 , g ∈ G}.
2.4 Clauses and dependencies 107
which simplifies to \
L(A ∪ {s}) = L(A).
s∈S
For any -irreducible L-closed set this is impossible. The -irreducible L-closed
T T
elements therefore all must be object intents of K. Since M is assumed to be finite,
the system of L-closed sets is generated by its irreducibles, and L is the set of
implications of the subcontext induced by these objects.
×
` [
a
Sm = Gm , {m} × Mm , ∇ ,
m∈M m∈M m∈M
with
(ga | a ∈ M ) ∇ (m, n) ⇔ gm Im n.
The formal contexts Sm (Gm , {m} × Mm , Im ) with
• •
•
gm Im (m, n) : ⇔ gm Im n
Proposition 32 Each context derived from a complete6 many-valued context via plain
scaling is isomorphic to a subcontext of the semiproduct of the scales used.
(m(g) | m ∈ M )
happens for no m ∈ M , then from each Gm an object vm can be selected such that
Am ⊆ v m 0
and S ∩ vm
0
= ∅. CombiningS these 0results in an object (vm | m ∈ M ) of
the semiproduct with object intent m∈M vm ,V
which contains A and is disjoint
from S, i.e., which does not respect the clause A → S.
W
The combination of the last two propositions with Theorem 9 now finally clari-
fies which rules apply to the implication inference of derived contexts: the Arm-
strong rules and all rules R∧A→∨S of factor scale clauses.
Example We illustrate the effect of these results with the “smallest non-trivial
example”. It lists the possible outcomes of a test that consists of two parts as
follows:
A natural interpretation is to interpret the attribute values pass and fail as nega-
tions of each other. This suggests scaling all three many-valued attributes with
the dichotomic scale. Object and attribute names of the abstract scale must there-
fore be adapted to this specific case. Plain conceptual scaling with these scales
leads to the one-valued context shown in Figure 2.20.
The logic of the implications of this example is obvious even without any the-
ory: The overall result is “pass” if and only if both parts are passed. Surprisingly,
however, the canonical basis of the derived context consists of eight implications,
not two.
6 The completeness assumption can be avoided by modifying the scales, see [138].
2.4 Clauses and dependencies 109
Figure 2.19 Abstract, concrete and factor scale (using “part 2” as an example).
The reason is that the background knowledge is not used for the computation
of the canonical base.
W The clauses of the (abstract) dichotomic scale are ∅ →
V
0 ∨ 1 and 0 ∧ 1 → ∅. >
That is to say, if two objects have the same values with respect to all attributes
from X the same must be true for the attributes from Y . The term “functional”
can be explained as follows: Y is functionally dependent on X if and only if there
is a map f : W X → W Y with
KO := (G × G, M, IO )
KN := (P2 (G), M, IN )
2.4 Clauses and dependencies 111
by
{g, h} IN m : ⇔ m(g) = m(h).
Then,
P2 (G) := {{g, h} | g, h ∈ G, g 6= h}.
The contexts defined in this way exactly fit the above-mentioned definitions
of the dependencies and it is easy to prove the following proposition:
Proposition 34 In (G, M, W, I) the attribute set Y is functionally dependent on X if
and only if the implication X → Y holds in the context KN . In (G, M, W, I) the attribute
set Y is ordinally dependent on X if and only if the implication X → Y holds in the
context KO .
Hereby we have traced back the theory of functional and ordinal dependen-
cies completely to the theory of implications. In particular, the algorithm men-
tioned in the previous section can also be used for the creation of a basis for the
functional or ordinal dependencies, respectively.
The translation works even if the many-valued context (G, M, W, I) is not com-
plete. In this connection, first of all we observe that Y is ordinally dependent on
X if and only if this is true for every single attribute in Y , i.e., if {n} is ordinally
dependent on X for every n ∈ Y . This means that it is sufficient to state in which
cases a single attribute is dependent on an attribute set. For the general case, this
can be formulated as follows:
Definition 49 (Def. 48 without the completeness assumption)
Let (G, M, W, I) be a many-valued context with an order relation ≤m on the set
m(G) of values for each attribute m ∈ M . If X ⊆ M is a set of attributes and
n ∈ M is an attribute, we say that n is ordinally dependent on X if
and n(g) 6≤ n(h) always implies that there exists an attribute m ∈ X with m(g) 6≤
m(h). ♦
In order to adapt Proposition 34, we have to modify the definitions of the
contexts KN and KO . We introduce a copy m̂ for every attribute m ∈ M which
is not complete. These new attributes have to be different from each other and
must not belong to M . We add the set
M̂ := {m̂ | dom(m) 6= G}
to the attribute set of the one-valued context. In the case of complete contexts,
we have M̂ = ∅, in general
with
{g, h} IN m̂ : ⇔ (g, h) IO m̂ : ⇔ g ∈ dom(m) and h ∈ dom(m)
112 2 Determination and representation
and, as above,
2.1
The algorithm in Theorem 5 has been taken from [143], see also [133]. For a dis-
cussion of other algorithms for finding formal concepts, see Section 2.4.4 of [154].
Further developments can be found in Ganter and Reuter [155], [135], Krolak–
Schwerdt, Orlik and Ganter [235]. With respect to complexity see Skorsky [344].
Schütt [339] gives an estimate of the number of concepts depending on |I|:
3 √|I|+1
|B(G, M, I)| ≤ ·2 −1 (for |I| > 2).
2
2.5 Notes, references, and trends 113
Other bounds are discussed in [1]. Finding the number of formal concepts is
known to be a #P -complete problem (Kuznetsov [245]).
2.2
The example of the waters from Figure 2.9 has been taken from the paper [220] by
Kipke and Wille. The automatic generation of diagrams has been discussed in de-
tail in the works of Skorsky, Luksch and Wille, see [344], [265] and [403] but also
Gepperth [167]. Besides, there are numerous implementations. Early ones were
Diagram for DOS by Frank Vogt and Conexp by Serhiy Yevtushenko. Toscana
[225] was a commercially available program system which facilitates and im-
proves the access to databases by means of elaborate nested line diagrams. See
also [407], [406] as well as Kühn & Ries [243]. The geometrical method has been
described in [400] and in [363] and has been supported by a computer program
by Kark [209]. Figure 2.5 is hand-drawn by R. Wille. Skorsky [343] has examined
the rule of parallelograms. Methods for automatic drawing were proposed, e.g.,
by Zschalig [430], in [139], and, more recently, by Dürrschnabel et al. [112].
[381] gives the “Basic Theorem” on line diagrams.
Other ways of representing contexts and concept lattices have been suggested,
which we shall not discuss here. See [400], Bokowski and Kollewe [57], Kollewe
[224], Lengnink [259], [260].
2.3
The results of Duquenne and Guigues permit a more effective algorithmic im-
plementation of the attribute exploration process, which had already been sug-
gested earlier by Wille. This has been described in [143], [133]. A remarkably
early application of this technique was realized by Reeg and Weiss [311]. In the
case of their investigation, the attribute set consisted of 50 common properties of
finite lattices.
Stumme [359] allows exceptions and background implications.
The method of attribute exploration has been further developed in different
ways (cf. [402]): On the one hand into concept exploration [397] (see also Klotz
and Mann [222]), which instead of attributes uses concepts. A specialization of
concept exploration to the distributive case which has practical applications to
knowledge acquisition is presented by Stumme [361].
On the other hand, the attribute exploration can be developed into rule ex-
ploration where the implications are replaced by Horn clauses from predicate
logic. This has been investigated by Zickwolff [429].
2.4
The first two subsections are based on [138] and [137], see also Ganter and
Krauße [150]. In his thesis, Krauße has computed clause bases for many stan-
dard scales.
Most of the results on dependencies have been taken from [159]. A uniform
theory of the dependency of many-valued attributes has been sketched in [398],
compare also [158]. The Θ-dependencies can be looked up in Stöhr and Wille
[353]. Umbreit [371] furthermore examines implications and dependencies be-
tween fuzzy attributes. See below for more recent contributions.
Research developments
There are numerous algorithms for computing formal concepts and also several
comparative studies, see e.g. Section 2.4.4 in [154]. That book also describes 30
other Concept Analysis algorithms in detail. The reason we limit ourselves here
to the Next Closure algorithm, which is fast but not the theoretically fastest, is its
captivating simplicity and the versatility it offers. A better worst case complexity
is provided by the algorithm of Nourine and Raynaud [293].
2.5 Notes, references, and trends 115
Implications
Database dependencies
If one wishes to examine parts of a rather complex concept system, it seems rea-
sonable to exclude some objects and/or attributes from the examination. We shall
describe the effects of this procedure on the concept lattice. The concept lattice
of a subcontext always has an order-embedding into that of the original context,
as will be shown in Section 3.1. Much more information can be obtained when
dealing with compatible subcontexts, which will be introduced later in this sec-
tion. It is easy to identify these particular subcontexts by means of the arrow
relations. One then obtains a factor lattice of the original concept lattice. The in-
terrelations between factor lattices, congruence relations and such subcontexts
will be described in the second section. Tolerance relations are generalized con-
gruence relations and also lead to a factor lattice. They correspond to certain
supersets of the incidence relation I, called block relations.
The complete sublattices of a concept lattice can also be described through parts
of the context, however not through subcontexts, but through certain subsets of
the incidence relation. Such closed relations will be defined in the third section.
In the fourth section we introduce bonds. These are relations that can connect
one formal context with a second one. They correspond in various ways to mor-
phisms between concept lattices and are used for this purpose in several places
in this book.
3.1 Subcontexts
1We write I ∩ H × N for I ∩ (H × N ). Instead of (H, N, I ∩ H×N ) some authors simply use
(H, N ) or [H, N ]. We write K1 ≤ K to express that K1 is a subcontext of K.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 117
B. Ganter, R Wille, Formal Concept Analysis, https://doi.org/10.1007/978-3-031-63422-2_4
118 3 Parts, factors, and bonds
We begin the section by asking how the concept system of a subcontext is related
to that of (G, M, I). If one merely omits attributes, that is, if one considers for a
subset N ⊆ M the context (G, N, I ∩ G × N ), then the answer is quite simple.
Any attribute extent of (G, N, I ∩ G × N ) is also an attribute extent of (G, M, I),
and since all concept extents are intersections of attribute extents, we obtain:
is a -preserving order-embedding.
W
× × × × ×
× × - × ×
× × × × ×
e e
@ @
e @e e u @e
@ @ - @ @
e @e @e e @e @e
HH
@
e @e
HH
An example is shown in Figure 3.1. If we combine the two parts of the propo-
sition, we obtain:
3.1 Subcontexts 119
(A, B) 7→ (B 0 , B 00 ).
These order-embeddings are bijective if (H,
V N, I ∩ H×N ) is a dense subcon-
text, i.e., if γH is -dense and dually µN is -dense in B(G, M, I).2
W
If ϕ : B(G, M, I) → (L, ≤) is an order-preserving mapping, then α := ϕ ◦ γ
and β := ϕ ◦ µ are maps α : G → L, β : M → L with
g I m ⇒ αg ≤ βm.
If, conversely, (α, β) is a pair of maps satisfying this condition, then, for instance,
the map _
ϕ(A, B) := αg
g∈A
3
is order-preserving. A useful special case is considered in the next proposition:
g I m ⇔ αg ≤ βm.
(H, N, I ∩ H×N )
2 A clarified doubly founded formal context has, of course, a unique smallest dense subcontext,
consisting of its irreducible objects and attributes.
3 See also Section 7.2.
120 3 Parts, factors, and bonds
can be expressed in terms of those of (G, M, I): If A ⊆ H, then the set of common
attributes with respect to (H, N, I ∩ H×N ) is equal to A0 ∩ N . Dually, the extent
of (H, N, I ∩ H×N ) belonging to a set B ⊆ N is equal to B 0 ∩ H. However,
the concepts of a subcontext cannot simply be derived from those of (G, M, I)
by restricting their extent and intent to a subcontext. This can be done only for
compatible subcontexts, which will be examined next.
Compatible subcontexts
↓ ↓ ↓
→× ×
→××× ×
→ × ×
×××
a b c d e c a b e d a b c d e .
.\ a b c d e
1 × . × .
% b b b b b 1 .
... .
. 1 × ×
2 × × × .
% × 2 .
. 2 × × × ×
b b b b
?
3 . × .
% × 3 .
..... 3 × ×
4 .
% .
% × × × 3 4 1 2 4 .
... 4 × × ×
the arrow-closed subcontexts in terms of the concepts of a context. For this pur-
pose we need the transitive closure of the arrow relations, as introduced in the
following definition:
Definition 53 For g ∈ G and m ∈ M we write g . . m if there are objects g =
g1 , g2 , . . . , gk ∈ G and attributes m1 , m2 , . . . , mk = m ∈ M with gi . mi for
i ∈ {1, . . . , k} and gj % mj−1 for j ∈ {2, . . . , k}. The complement (the negation)
of this relation is denoted by . \ , i.e., g .
. \ m ⇔ not g .
. . m. ♦
Proof From the presuppositions doubly founded and reduced follows by Proposi-
tion 15 that for every object g there is an attribute m with g %
. m, and dually.
First of all, let (H, N, I ∩ H×N ) be arrow-closed. If g ∈ H and g % . m, then
m ∈ N must be true, i.e., g ∈ H holds if and only if there is an n ∈ N with
g .. n. Consequently, g ∈ G \ H if and only if g . \ n for all n ∈ N , i.e.,
.
G \ H = N .. \
. Now assume that m ∈ M \ N and g % . m. g ∈ H is impossible
because (H, N, I ∩ H×N ) is arrow-closed. Therefore, we get g ∈ G \ H, g ..m
3.1 Subcontexts 123
the conditions of doubly foundedness is proved thereby, the other one follows
dually.
We should also mention that dense subcontexts are always compatible:
Proof 1) ⇔ 3): (H, N, I ∩ H×N ) is denseWif and only if both, for every object
g ∈ G there is a subset X ⊆ H with γg = x∈X γx, i.e., with g 0 = X 0 and thus
g ∈ X 00 , and the dual condition holds for the attributes. Because of γx ≤ γg for
all x ∈ X we have X ⊆ g 00 and thus the condition from 3) for the case A = g 00 .
The more general condition follows without difficulty.
3) ⇒2): In order to show that (H, N, I ∩ H×N ) is compatible, we prove the
conditions b) from Proposition 41: If A ⊆ G, then A0 is an intent and on account
of 3) it satisfies (A0 ∩ N )00 = A0 , which yields (A0 ∩ N )0 = A00 and thus b1). b2)
is dual. The injectivity of ΠH,N immediately follows from (A ∩ H)00 = A.
2) ⇒ 3): If ΠH,N is injective and (A, B) is a concept of (G, M, I), then (A ∩ H)00 =
A must hold, otherwise (A ∩ H)00 and A would be different extents having the
same intersection with H.
In the preceding section we have seen that, for a compatible subcontext (H, N, I ∩
H×N ) of (G, M, I), the map ΠH,N is a surjective complete homomorphism of
B(G, M, I) onto B(H, N, I∩ H×N ), i.e., that the concept lattice of the subcontext
is always a homomorphic image of B(G, M, I). We shall now examine whether
the converse is true as well, i.e., whether every surjective complete homomor-
phism, every homomorphic image can be described in terms of a subcontext. In
the case of finite contexts this is true, in the case of infinite contexts not in general.
In order to clarify the situation, we require a notion from lattice theory, namely
that of the complete congruence relation.
Definition 54 A complete congruence relation of a complete lattice (L, ≤) is an
equivalence relation Θ on L satisfying:
^ ^ _ _
xt Θ yt for t ∈ T ⇒ ( xt ) Θ ( yt ) and ( xt ) Θ ( yt ).
t∈T t∈T t∈T t∈T
3.2 Complete congruences and tolerances 125
We define
[x]Θ := {y ∈ L | x Θ y},
which is the equivalence class of Θ containing x. The factor lattice
is ordered as follows:
In order to demonstrate that this is really an order relation we can, for instance,
argue as follows: If we define
^ ^
xΘ := {y ∈ (L, ≤) | y Θ x} = [x]Θ
_ _
and xΘ := {y ∈ (L, ≤) | y Θ x} = [x]Θ
for x ∈ (L, ≤), this immediately yields [x]Θ = xΘ , xΘ and [x]Θ ≤ [y]Θ ⇔ xΘ ≤
e
@
e @e e
@ @ @
e @e @e e @e
@ @ @
@ @ @e
@ @
@e
@ @e
@ e
@
@e
It asserts among other things that every homomorphic image of a complete lat-
tice can be found within the lattice itself, namely as a factor lattice.
Theorem 10 (Homomorphism Theorem) If Θ is a complete congruence relation of
a complete lattice (L, ≤), then x 7→ [x]Θ is a complete homomorphism of (L, ≤) onto
(L, ≤)/Θ. If, conversely, ϕ : (L1 , ≤) −→ (L2 , ≤) is a surjective complete homomor-
phism between complete lattices, then
and thus y ∈ [x]Θ. This means that [x]Θ = xΘ , xΘ . The maps x 7→ xΘ and
B(H, N, I ∩ H×N ) ∼
= B(G, M, I)/ΘH,N ,
with
It is easy to identify the smallest and the largest elements of the congruence
classes. If (A, B) is a concept, the smallest element of the congruence class
[(A, B)]ΘH,N is the concept ((A ∩ H)00 , (A ∩ H)0 ) and the largest is the concept
((B ∩ N )0 , (B ∩ N )00 ).
We say that a complete congruence Θ is induced by a subcontext if there is a
compatible subcontext (H, N, I ∩ H × N ) with Θ = ΘH,N . Using the connection
between compatible subcontexts and congruences, we shall prove the following:
in the case of a doubly founded concept lattice every congruence is induced by
128 3 Parts, factors, and bonds
(GΘ , MΘ , I ∩ GΘ × MΘ ).
g ∈ GΘ ⇔ there is X ⊆ H with X 0 = g 0 .
(GΘ , MΘ , I ∩ GΘ × MΘ ).
Now we turn to the second part of the question: Which congruences are in-
duced by subcontexts? On account of the propositions we know that H = GΘ
and N = MΘ can be chosen, if Θ is at all induced by a subcontext. It is easy
to state congruences which do not have this form, these examples are however
infinite. The following propositions provide an exact clarification.
B(H, N, I ∩ H×N ) ∼
= B(G, M, I)/Θ,
and the isomorphism (A, B) 7→ [(A, B)]Θ maps the supremum-dense set {γh |
h ∈ H} onto {[γh]Θ | h ∈ H}, i.e., this set is supremum-dense in B(G, M, I)/Θ,
and dually {[µn]Θ | n ∈ N } is infimum-dense. Because of H ⊆ GΘ and N ⊆ MΘ
thus the direction “⇒” of the assertion follows.
We begin the proof of the other direction by showing that (GΘ , MΘ , I ∩ GΘ ×
MΘ ) is compatible under the conditions specified. Assume that h ∈ GΘ and that
m ∈ M with (h, m) ∈ / I. Then [γh]Θ 6≤ [µm]Θ and, since {[µn]Θ | n ∈ MΘ } is
infimum-dense in B(G, M, I)/Θ, there is some n ∈ MΘ \ h0 with µn ≥ µm, i.e.,
n0 ⊇ m0 . This yields 41.a1) and dually 41.a2).
In order to show that Θ is induced by (GΘ , MΘ , I ∩ GΘ × MΘ ), we have to prove
that
(A, B) Θ (C, D) ⇔ A ∩ GΘ = C ∩ GΘ
for (A, B), (C, D) ∈ B(G, M, I). Let (A, B) be the smallest concept in the Θ-class
containing (A, B). For h ∈ GΘ we have γh ≤ (A, B)W⇔ γh ≤ (A, B), since γh also
is the smallest element of a Θ-class, and (A, B) = {γh | h ∈ A ∩ GΘ } because
{[γh]Θ | h ∈ GΘ } is supremum-dense. (A, B) and (C, D) are congruent if and
only if the classes [(A, B)]Θ and [(C, D)]Θ have the same smallest element, i.e.,
if A ∩ GΘ = C ∩ GΘ .
If [v]Θ is -irreducible in B(G,
W M, I)/Θ, then the smallest element of the con-
W
gruence class [v]Θ must also be -irreducible and thus must be an W object concept
γg with g ∈ GΘ . Hence, the set {[γh]Θ | h ∈ GΘ } contains all -irreducible el-
ements, and likewise {[µn]Θ | n ∈ MΘ } contains all -irreducible elements of
V
B(G, M, I)/Θ. Thus, from Proposition 48 we can infer:
130 3 Parts, factors, and bonds
B ∩ MΘ = D ∩ M Θ ⇒ B ∩ MΨ = D ∩ MΨ
for all (A, B), (C, D) ∈ L
⇔ GΨ ⊆ GΘ and MΨ ⊆ MΘ .
(H1 , N1 , I ∩ H1 × N1 ) ≤ (H2 , N2 , I ∩ H2 × N2 )
:⇔ H1 ⊆ H2 and N1 ⊆ N2 ,
under the preconditions specified, the ordered set of the arrow-closed subcon-
texts is dually isomorphic to the lattice of complete congruences. Now, how-
ever, the union as well as the intersection of arrow-closed subcontexts are arrow-
closed too. Therefore, the lattice of arrow-closed subcontexts is completely dis-
tributive. Thus, Proposition 43 makes it easy to state a context for the congruence
lattice as well.
Theorem 13 The congruence lattice of a doubly founded concept lattice B(G, M, I) is
isomorphic to the completely distributive lattice B(G, M, .
.
\ ).
Proof If (G, M, I) is reduced, then every congruence is induced by exactly one
subcontext (H, N, I ∩ H×N ). Furthermore we know by Proposition 43 that those
subcontexts correspond to the concepts of (G, M, . .\ ): (H, N, I ∩ H×N ) induces
a congruence if and only if (G \ H, N ) is such a concept.
The order of those subcontexts is dual to that of the concepts of (G, M, . .
\)
as well as to that of the congruences. This means that the latter two must be
isomorphic to each other.
For the structure of B(G, M, . .\ ), however, it is irrelevant whether (G, M, I) is
reduced, provided that B(G, M, I) is doubly founded. In this case, we can switch
to the reduced context (Girr , Mirr , I ∩ Girr × Mirr ) with Girr and Mirr being the
set of irreducible objects and attributes, respectively. The . .-relation is inherited
by this subcontext, since in Definition 53, apart from g and m, there only appear
irreducible objects and attributes. Therefore, B(G, M, . .\ ) and B(Girr , Mirr , .
.
\)
are isomorphic, every concept of (G, M, . \ ) is of the form
.
(G \ Girr ) ∪ A, B ∪ (A..\
∩ (M \ Mirr )) ,
f
@
d f fc @ fe
@
@ @
f f @f
@
@ @
1 @ 3 @
@v
@ @ fa, b
@
4 @ 2
@f
@
Figure 3.5 The congruence lattice of the lattice from Figure 3.4 is at the same time the
lattice of arrow-closed subcontexts of Figure 3.3. The marked element corresponds to the
congruence from Figure 3.4 and the compatible subcontext in Figure 3.2.
^ ^ _ _
xt Θ yt for t ∈ T ⇒ ( xt ) Θ ( yt ) and ( xt ) Θ ( yt ).
t∈T t∈T t∈T t∈T
Θ . The dual definition [a]Θ := [(aΘ )Θ , aΘ ] also yields the blocks of Θ: Because of
((aΘ )Θ )Θ = aΘ and ((aΘ )Θ )Θ = aΘ we obtain [a]Θ = [aΘ ]Θ as well as [a]Θ = [aΘ ]Θ .
From aΘb and a ≤ b it follows that bΘ ≤ a ≤ b, i.e., a ∈ [b]Θ . Correspondingly,
from aΘb and a ≥ b it always follows that a ∈ [b]Θ . The blocks of a tolerance rela-
tion do not have to be disjoint, unless we are dealing with a congruence relation.
We have
[a]Θ ∩ [b]Θ 6= ∅ ⇔ aΘ ≤ (bΘ )Θ and bΘ ≤ (aΘ )Θ .
3.2 Complete congruences and tolerances 133
f f
@ @
@ @
f f
f @f f @ f
@ @@ @
@ @ @ @
f f @f f @f f f @ f f @f
@ @
@
@
@ @ @ @
@f @f @f @f
@ @
@ @
@f @f
Figure 3.6 The pairs which are linked together in the figure on the right (including the
pairs of neighboring elements) form part of a tolerance relation of the lattice on the left.
and correspondingly
^ _d ^
ψ( xt ) = ψ( xt ) = ψ(xt ).
' $
f
@
' f$
'
@
$
f @f
@ C@
' C$
@ C @
f f @f f @f
& %
@ C @
@ C @
@C f @f
& &% %
@
@
@f
& %
Figure 3.7 The blocks of the tolerance relation from Figure 3.6.
^ ^ _ _
B1 ≤ B2 : ⇔ B1 ≤ B2 (⇔ B1 ≤ B2 ). ♦
The definition says that the smallest elements of the blocks are ordered in
the same way as their largest elements. This is correct because of xΘ ≤ yΘ ⇔
(xΘ )Θ ≤ (yΘ )Θ . In fact, even more is true: The set of the upper bounds of the
blocks is closed under infima, that of the lower bounds is closed under suprema,
in analogy to the case of the complete congruences (cf. Theorem 11, p. 126). This
is described by the following theorem:
Theorem 14 With the order described above, L/Θ is a complete lattice (the factor lat-
tice of L by Θ). The following equations hold for blocks Bt and for elements xt , t ∈ T ,
of L respectively:
_ _^ ^ ^_
Bt = [ Bt ]Θ resp. Bt = [ Bt ] Θ
t∈T t∈T t∈T t∈T
_ _ ^ ^
[xt ]Θ = [ xt ]Θ resp. Θ
[xt ] = [ xt ] Θ
Proof The proofs of the equations follow easily from Proposition 53.
How can we describe complete tolerance relations of concept lattices in terms
of the contexts?
Definition 59 By a block relation of a context (G, M, I) we mean a relation J ⊆
G × M which satisfies the following conditions:
1. I ⊆ J,
2. for every object g ∈ G, g J is an intent of (G, M, I),
3.2 Complete congruences and tolerances 135
f
@
@
c f fd @ fe
@ 7 @
@ fg
@ @
a f b f ff
@f a b c d e f g
1 × • × • •
1 @ 2 5 6
@ 2 • × × • •
@ @ 3 × × × × × • •
@f @f 4 • • × × × × ×
3 @ 4 5 • • × × •
@ 6 • • × • ×
@f 7 • × •
Figure 3.8 The block relation J belonging to the tolerance from Figure 3.6 additionally
contains the pairs marked by dots.
Theorem 15 The lattice of all block relations of (G, M, I) is isomorphic to the lattice
of all complete tolerance relations of B(G, M, I). The map β assigning to any complete
tolerance relation Θ the block relation defined by
is an isomorphism. Conversely,
g J = {m ∈ M | γg Θ (γg ∧ µm)}.
We claim that this is an intent of (G, M, I). For this purpose we consider the
concept
136 3 Parts, factors, and bonds
^
{µm | γg Θ (γg ∧ µm)} = (g JI , g JII ).
If n is an attribute of this concept, we get
^
µn ≥ {µm | γg Θ (γg ∧ µm)},
If we are aware that this infimum is in a Θ-relation with γg, we recognize that
(γg ∧ µn) Θ γg, i.e., that n ∈ g J . Hence, g J = g JII is an intent. Dually we prove
that every set of the form mJ is an extent of (G, M, I).
Now we start from a block relation J and define a relation τ (J) on B(G, M, I)
by
(A, B) τ (J) (C, D) : ⇔ A × D ∪ B × C ⊆ J.
Evidently τ (J) is reflexive and symmetric, and, if T is an index set and there are
concepts with (At , Bt ) Θ (Ct , Dt ) for t ∈ T , we argue as follows: For g ∈
T At we
have g J ⊇ Dt = T CtI and consequently g JI ⊆ CtII = Ct . Hence, for g ∈ t∈T At
we obtain g JI ⊆ t∈T Ct , i.e.,
!I
\
J JII
g =g ⊇ Ct ,
t∈T
which proves
!I
\ \
At × Ct ⊆ J.
t∈T t∈T
The dual argument proves, moreover, that τ (J) is compatible with suprema, i.e.,
that it is a complete tolerance relation.
Both maps β and τ are evidently order-preserving. In order to prove the the-
orem we furthermore have to show that they are inverse to each other. Let Θ be
a complete tolerance relation of B(G, M, I). We want to show that
3.2 Complete congruences and tolerances 137
According to Proposition 51, we may limit ourselves to the special case (A, B) >
(C, D). We have
For the last part of the proof let J be a block relation of (G, M, I). Then
(g, m) ∈ J ⇔ g ∈ mJ
⇔ g II ⊆ g JJ ∩ mJ and
(g II ∩ mI )I = (g I ∪ mII )II ⊆ (g J ∪ mJJ )JJ = (g JJ ∩ mJ )J
⇔ g II × (g II ∩ mI )I ⊆ J
⇔ (γg) τ (J) (γg ∧ µm)
⇔ (g, m) ∈ β(τ (J)).
B(G, M, I)/Θ ∼
= B(G, M, J).
2. The map I
(B , B), (C, C I ) 7→ (C, B)
is an isomorphism of the lattice of the blocks of Θ onto the concept lattice of (G, M, J).
3. If (C, B) is a concept of (G, M, J), then
I
(B , B), (C, C I ) = B(C, B, I ∩ C × B)
(X, Y )Θ = (X JI , X J ) and
Θ
(X, Y ) = (Y , YJ JI
) and consequently
138 3 Parts, factors, and bonds
((X, Y )Θ )Θ = (X JJ , X JJI ).
c, d, e
g
7@ c d e f g
a, b @ gf, g
g
@ 3×××
1,2 @ 5,6 4×××××
5 ××
@g
@ 6 × ×
3,4
Figure 3.9 The concept lattice B(G, M, J) of the block relation J is isomorphic to the fac-
tor lattice by the tolerance relation. As an example, we state the subcontext belonging to
the concept ({3, 4, 5, 6}, {c, d, e, f, g}) of J . Its concept lattice is isomorphic to the corre-
sponding block of the tolerance.
We close this section with two observations. The first deals with the question
which families of intervals form the systems of the blocks of a tolerance. This is
answered by the following theorem:
Theorem 16 Let L be a complete lattice, T an index set and
F := {[xt , xt ] | t ∈ T }
3. There is an order ≤ on T with respect to which (T, ≤) is a complete lattice and there
are maps
α : (T, ≤) → L, injective and -preserving,
W
i.e., the result is again an upper bound of a block. This proves a) and dually we
infer b). c) again follows from Proposition 53: xΘ ≤ y Θ ⇔ (xΘ )Θ ≤ (y Θ )Θ .
2 ⇒ 3: We order T by s ≤ t : ⇔ xs ≤ xt . Because of c) this is equivalent to
xs ≤ xt . Hence, the map defined by α(t) := xt is an order-isomorphism of (T, ≤)
on {xt | t ∈ T }, correspondingly α(t) := xt defines an order-isomorphism of
(T, ≤) on {xt | t ∈ T }. Therefore, according to a) and b), (T, ≤) is a complete
lattice.
3 ⇒ 1: We first show that under the conditions mentioned under 3, the set
[
U := {[α(t), α(t)] | t ∈ T }
i.e., _ _
( {xs | s ∈ S}, {ys | s ∈ S}) ∈ Θ.
Hence, Θ is -compatible (and dually, of course, -compatible as well), i.e., it is
W V
a complete tolerance.
Finally we have to show that the intervals [α(t), α(t)] are in fact the blocks of Θ,
i.e., the maximal sets of elements which are pairwise related under the relation
Θ. If [u, v] is a block, then (u, v) ∈ Θ, i.e., u, v ∈ [α(t), α(t)] for some t, hence every
140 3 Parts, factors, and bonds
block of Θ is of this form. On the other hand, every [α(t), α(t)] is maximal too;
from [α(s), α(s)] ⊆ [α(t), α(t)] we can infer α(s) ≤ α(t), that is s ≤ t, as well as
α(t) ≤ α(s), that is t ≤ s. Together this yields s = t.
The second observation establishes a link between compatible subcontexts and
block relations. Transitive tolerance relations are congruence relations, hence we
can describe a complete congruence relation (under suitable conditions, cf. The-
orem 12) in two ways: in terms of a compatible subcontext and in terms of a block
relation.
Proposition 54 Let Θ be a complete congruence relation of a doubly founded concept
lattice B(G, M, I), (GΘ , MΘ , I ∩ GΘ × MΘ ) the corresponding saturated compatible
subcontext and J = β(Θ) the block relation for Θ. Then:
(g, m) ∈ J ⇔ g 00 ∩ GΘ ⊆ m0
⇔ m00 ∩ MΘ ⊆ g 0
GΘ = {g ∈ G | g I = g J }
MΘ = {m ∈ M | mI = mJ }.
We have to show that J is a closed relation with U = B(G, M, J). It is evident that
U ⊆ B(G, M, J). Thus, it remains to be shown that every concept of (G, M, J)
belongs
T to U . We first prove this for the object concepts: Assume that g ∈ G and
D := {A | (A, B) ∈ U, g ∈ A}. D is an extent of (G, M, J), and consequently
g JJ ⊆ D. For every attribute m ∈ g J there exists a concept (A, B) ∈ U with
(g, m) ∈ A×B and because of D ⊆ A it follows that m ∈ DJ . Therefore, g J = DJ
and g JJ = D. This shows that for every g ∈ G the concept (g JJ , g J ) belongs to U .
Every concept of B(G, M, J) is however the supremum of such object concepts,
thus U ⊇ B(G, M, J), which remained to be proved.
This means that the closed relations are in a 1-1 correspondence to the com-
plete sublattices. The map
[
C(U ) := {A × B | (A, B) ∈ U }
maps the set of complete sublattices bijectively onto the map of closed relations
of B(G, M, I).S It is furthermore order-preserving, U1 ⊆ U2 ⇔ J1 ⊆ J2 . However,
C is neither – nor –preserving. The intersection of closed relations does not
T
necessarily have to be closed, the closed relations in general do not form a closure
system. This is surprising in so far as the family of complete sublattices does
form a closure system: for every subset T of a complete lattice, the intersection
of all lattices containing T is also a sublattice (namely the complete sublattice
generated by T ).
If F is a family of closed relations and D := F, then there is, nonetheless,
T
always a largest closed relation in D, namely
[
J := {A × B | (A, B) concept, A × B ⊆ D}.
form precisely the intersection of those sublattices, i.e., they themselves are a
complete sublattice. These considerations yield the following proposition:
Proposition 55 For every set T ⊆ B(G, M, I) of concepts, there is a smallest closed
relation J of (G, M, I) containing all sets A × B with (A, B) ∈ T . B(G, M, J) is the
complete sublattice of B(G, M, I) generated by T .
a b c d e f g h
1 × ×
2×××
3××⊗ ⊗
4 ⊗⊗⊗
5 ⊗⊗⊗⊗⊗
6 ⊗⊗ ⊗ ⊗
7 × ××
8 ××××
9 ⊗⊗⊗⊗⊗⊗
10 × × ×
X JJ ⊇ X JI
Proposition 57 The closed relations of a context (G, M, I) are precisely those subrela-
tions J ⊆ I which satisfy the following condition:
(C) (g, m) ∈ I \ J implies (h, m) 6∈ I for some h ∈ G with g J ⊆ hJ as well as (g, n) 6∈ I
for some n ∈ M with mJ ⊆ nJ .
u
H g
e e
@He Hf
d e @H H ec
H H H
e @H e HH
HH HH HHH HH
HuH @eH
@ H HH H HH
H HHH
H b
H H
@e H Hu H
uHHHe7HHeH
@ @ HHH@ HHH
H
H 4
8 HH@ HHHHu H
@H
H eH
H u H10 H ea
H
Hu
H@ H@ H H
H
@ H H
6 H
5 H He 2
H
HH H
h 3
eP Hu
H
1 PP 9
PP
Pu
PP
Figure 3.11 Diagram of the concept lattice for the context from Figure 3.10. The sublattice
consisting of the blackened elements, belongs to the above-mentioned closed relation.
(H, N, I ∩ H × N )
In the following proposition this refers to the arrow relations in the context
(G, M, J):
Proposition 59 Let (G, M, J) be a doubly founded clarified context. Then the following
statement holds: J is a closed relation of (G, M, I) if and only if
J ⊆ I ⊆ G × M \ (. ∪ %).
Intervals
Full rows and full columns of a context belong to every closed relation and it is
sometimes awkward to have to carry them along. For simplification purposes,
we therefore occasionally use the notation
:= M 0 × M ∪ G × G0 .
I ∩ A × M, I ∩ G × D, I ∩ (A × M ∪ G × D)
B0
C
C0
Figure 3.12 With reference to Proposition 61: Between (B 0 , B) and (C, C 0 ) lie precisely
the concepts of the context (C, B, I ∩ C × B).
It turns out that in this subcontext the objects 1, 2, 3, 5, 9 and 10 as well as the
attribute c are reducible. The reduced context is presented in Figure 3.13 together
with its concept lattice. It is an interval in the center of Figure 3.11.
c
d e f g
4 × × e c@ cg
@
6 × × dc @@c @ @ cf
7 × × 4 HH 6 7
H c
Further examples of closed relations are obtained from formal concepts in-
variant under an automorphism group and from dismantling:
If Γ is a group of automorphisms of the context (G, M, I), i.e., of pairs of maps
(α, β) with
α : G → G, β : M → M, g I m ⇔ α(g) I β(m),
Dismantling
It may happen that a closed relation differs very little from the incidence relation
I, in the extreme case only by one “cross”. This case corresponds to the disman-
tling of doubly irreducible elements. Therefore, we shall give a short description,
only sketching the order-theoretic results and referring to the corresponding lit-
erature for the proofs.
An element a of an ordered set shall be called doubly irreducible if a has ex-
actly one lower neighbor a∗ and exactly one upper neighbor a∗ and furthermore
the conditions
x < a ⇒ x ≤ a∗ , x > a ⇒ x ≥ a∗
3.4 Bonds and connections 147
We call bonds those relations between formal contexts that generate adjunctions
between the associated concept lattices. This connection is made precise by
Proposition 66. But before we clarify some basic definitions and facts.
♦
This is obviously equivalent to:
• Every extent of (G, N, R) is an extent of (G, M, I) and
• every intent of (G, N, R) is an intent of (H, N, J).
Since intersecting extents yields extents and intersecting intents gives intents,
the intersection of bonds is always a bond. From this follows the next proposi-
tion:
Proposition 64 The set of all bonds from (G, M, I) to (H, N, J) is a closure system on
G × N.
The corresponding closure operator is easy to describe, since a subset of G ×
N is closed if and only if there is an intent of (H, N, J) in each “row” and an
extent of (G, M, I) in each “column”. Thus, the extent closure operator of the
first context is to be applied column-wise and the intent closure operator of the
second context is to be applied row-wise. This usually has to be repeated several
times alternately.
Bonds were introduced by R. Wille [393] to describe subdirect products of
concept lattices, see Chapter 5 for an introduction. Subdirect products are certain
complete sublattices of the direct product, and since direct products correspond
to context sums and complete sublattices to closed relations, the following is natural
(and an instance of Theorem 37 (p. 218)):
Proposition 65 R ⊆ G × N is a bond from (G, M, I)
to (H, N, J) if and only if M N
C := I ∪ R ∪ J ∪ (H × M ) G I R
is a closed relation of the context sum
(G, M, I) + (H, N, J) H @ J
@
(assuming G ∩ H = ∅ = M ∩ N ).
3.4 Bonds and connections 149
Proof The formal concepts of the sum (cf. Definition 30) are precisely the pairs
of the form
(A1 ∪ A2 , B1 ∪ B2 ),
where (A1 , B1 ) ∈ B(G, M, I) and (A2 , B2 ) ∈ B(H, N, J). Now let (U, V ) be a
concept of (G, N, R). Then (U ∪ V J , V ∪ U J ) is a concept of
(G ∪ H, M ∪ N, C).
(G ∪ H, M ∪ N, C),
U = (B ∩ M )I ∩ V R ,
Adjunctions between concept lattices were introduced on Page 19, see also
Proposition 11 there. They are in 1-1-correspondence to bonds, as the following
result shows.
Proposition 66 From every bond R ⊆ G × N between formal contexts (G, M, I) and
(H, N, J) we obtain an adjunction, i.e., a pair (ϕR , ψR ) of mappings
ϕR (A, AI ) := (ARJ , AR ), ψR (B J , B) := (B R , B RI ).
Each adjunction is obtained in this way from exactly one bond, in fact,
For the proof see that of Theorem 20 below, where Corollary 2 states just the dual
of Proposition 66.
150 3 Parts, factors, and bonds
Adjunctions differ from Galois connections only in that the connecting map-
pings are order-preserving instead of order-reversing. This has an important con-
sequence: adjunctions can be concatenated.
The category of concept lattices with adjunctions as morphisms has remark-
able properties (∗ -autonomous, see Mori [288], Borgwardt [64]). Since (as stated
in Proposition 66) such adjunctions are represented by bonds, it is natural to ask
for a concatenation of bonds. However, the usual relation product does not come
into question for this, simply because bonds do not fit together appropriately;
they always go from objects to attributes. The following definition, on the other
hand, will work.
Definition 62 Let Ki := (Gi , Mi , Ii ) (i ∈ {1, 2, 3}) be formal contexts and let
J12 ⊆ G1 × M2 and J23 ⊆ G2 × M3 be relations. We define
♦
Proposition 67 (cf. Figure 3.14) If J12 is a bond, then
[
J12 ◦ J23 = B J12 × AJ23 ,
(A,B)∈B(K2 )
Proof If (A, B) ∈ B(K2 ) and g ∈ B J12 then B ⊆ g J12 and thus A ⊇ g J12 I2 . Any
m ∈ AJ23 therefore satisfies g J12 I2 ⊆ mJ23 , which proves that B J12 × AJ23 ⊆
J12 ◦ J23 . For the other direction let g ∈ G1 , A := g J12 I2 and B := AI2 . Then
(A, B) ∈ B(K2 ), and, since g J12 is an intent of K2 , B = AI2 = g J12 I2 I2 = g J12 ,
which implies g ∈ B J12 . If mJ23 contains g J12 I2 for some m ∈ M3 , then mJ23
contains A (= g J12 I2 ) and thus m ∈ AJ23 .
B J12
J12 J12 ◦ J23
I2 J23 A
B AJ23
M1 M2 M3 M1 M2 M3
g
G1 I1 J12 G1
G2
n
G2 I2 J23 g J12 I2
mJ23
|{z}
G3 I3 g J12 G3
|{z} m
g J12 I2 J23
simply reflects the fact that m belongs to the J23 -intent of g J12 I2 iff the J23 -extent
of m contains g J12 I2 .
which by Proposition 68 this is equal to g J12 I2 J23 , the latter being an intent of K3 .
Dually one shows that for fixed m ∈ M3
152 3 Parts, factors, and bonds
I1 J12 J13
I2 J23
I3
if and only if
J13 = J12 ◦ J23 .
Proof (g, m) is an element of the bond corresponding to ϕ23 ◦ϕ12 iff ϕ23 ◦ϕ12 γg ≤
µm. The intent of ϕ23 ◦ϕ12 γg is g J12 I2 J23 . It was shown in the proof of Theorem 18
that this is equal to
{m ∈ M3 | (g, m) ∈ J12 ◦ J23 }.
So J13 is indeed the bond corresponding to ϕ23 ◦ ϕ12 and ψ12 ◦ ψ23 .
Remark: The results outlined here can be expressed particularly clearly in the
language of mathematical Category Theory. The formal contexts as objects form
a category with the bonds as morphisms. This category is naturally equivalent
to the category of complete lattices with the residuated mappings and there-
fore inherits the special properties of that category (see Theorem 3.8 of Borg-
wardt [64]). The functors that establish this equivalence are K 7→ B(K) on the
one hand and (L, ≤) 7→ (L, L, ≤) on the other, with the actions on the morphisms
explained above.
5 Here ϕij is an abbreviation for ϕJij and ψij abbreviates ψJij .
3.4 Bonds and connections 153
The identical mapping is residuated for each concept lattice B(G, M, I). The
corresponding bond is the incidence relation I. Other self-bonds are, e.g., the
closed relations and the block relations (see Definitions 60 and 59).
The basic construction for Formal Concept Analysis was described in Section 1.1:
From a relation I ⊆ G × M between two sets G and M a Galois connection be-
tween the powerset lattices P(G) and P(M ) is obtained by the derivation oper-
ators, and from it the concept lattice B(G, M, I).
Galois connections between complete lattices were introduced in Section 0.4
and there it was already mentioned that the definition can be extended to arbi-
trary ordered sets without modifications. Now we generalize even further and
introduce Galois connections between formal contexts. Here the connection is
established by a pair of relations rather than mappings.7 In the first condition of
Definition 63, which is illustrated in Fig. 3.17, g I hΨ is short for hΨ ⊆ g I , etc. The
second condition will be simplified in Proposition 71.
Definition 63 A (relational) Galois connection between (G, M, I) and (H, N, J)
is a pair (Φ, Ψ ) of relations
Φ⊆G×N and Ψ ⊆H ×M
satisfying
1. g I hΨ ⇔ h J g Φ for all g ∈ G, h ∈ H (the Galois condition),
2. Φ is the largest relation fulfilling the Galois condition for the given Ψ , and
conversely. ♦
6 i.e., ∀x ϕJ (x) ≤ x
7 See Definition 103 on p. 296 for the latter.
154 3 Parts, factors, and bonds
M N
G I Φ
H Ψ J
Figure 3.17 The Galois condition requires that the left hand part hΨ of a row from the
lower part is contained in a row g I of the upper part iff the right hand side g Φ of the latter
is contained in the former, in hJ .
g ΦJ ⊆ {h ∈ H | g I hΨ },
which implies
g Φ ⊆ {h ∈ H | g I hΨ }J = g ΦJJ .
Thus from Φ we obtain another relation satisfying the Galois condition for Ψ by
replacing each g Φ by its closure g ΦJJ . This then is the largest possible choice. The
dual argument works for Ψ .
This proof also shows that one gets condition 2) of Definition 63 for free, so to
speak. If a pair of relations satisfies the Galois condition, then these relations can
be uniquely enlarged so that condition 2) is also satisfied. One only has to replace
g Φ and hΨ by the concept intents they generate.
The next proposition shows that Definition 63 in fact generalizes Definition 18.
p I q Ψ ⇔ p ≤1 ψ(q) ⇔ q ≤2 ϕ(q) ⇔ q J pΦ .
Now that the definition of a Galois connection has been generalized in several
steps (from complete lattices to arbitrary ordered sets and further to arbitrary
binary relations, from mapping pairs to relations), the next theorem shows that
we finally reach the starting point again.
Theorem 19 Whenever (ϕ, ψ) is a Galois connection between concept lattices
B(G, M, I) and B(H, N, J), then
g Φ = Y : ⇔ ϕ(g II , g I ) = (Y J , Y )
and
hΨ = X : ⇔ ψ(hJJ , hJ ) = (X I , X)
defines a relational Galois connection between (G, M, I) and (H, N, J).
Conversely we obtain from each relational Galois connection (Φ, Ψ ) between (G, M, I)
and (H, N, J) a Galois connection (ϕ, ψ) between the concept lattices B(G, M, I) and
B(H, N, J) by
Dual bonds
Theorem 20 From each relational Galois connection (Φ, Ψ ) between (G, M, I) and
(H, N, J) we obtain a dual bond R ⊆ G × H by means of
(g, h) ∈ R :⇔ g I hΨ ( ⇔ h J g Φ ).
g Φ := g RJ for g ∈ G,
hΨ := hRI for h ∈ H
defines a relational Galois connection. The two constructions are inverse to each other.
Proof If R is defined as above from a relational Galois connection (Φ, Ψ ) we get
g R = g ΦJ and hR = hΨ I for g ∈ G, h ∈ H.
ϕ(X, X I ) := (X R , X RJ ), ψ(Y, Y J ) := (Y R , Y RI )
defines a Galois connection between B(G, M, I) and B(H, N, J). Conversely, for every
Galois connection (ϕ, ψ),
A 7→ ARR for A ⊆ G
C 7→ C RR for C ⊆ H
K1 × K2 := (G1 × G2 , M1 × M2 , ∇)
with
(g1 , g2 ) ∇ (m1 , m2 ) :⇔ g1 I1 m1 or g2 I2 m2 .
If A has only one element, i.e., if A = {g}, we omit the brackets and write R(g)
instead of R({g}). Note that
[
R(A) = R(g).
g∈A
A proof of this theorem will be given below, but it requires some preparation.
Recall that the object set of the direct product (G, M, I) × (H, N, J) is G × H.
Subsets of this object set are the relations R ⊆ G×H, and applying the derivation
operator of the direct product leads to
I
g∈mr
I
m∈gr
equality of the first and the second set. The equality of the second and the third
is immediate from the definition of the derivation operators.
To verify the second part of the assertion, note that R∇ is a relation between
the object sets of the dual contexts. The first part, which has already been proven,
can be applied to this and confirms the assertion.
Proof of Theorem 21. We show more than is claimed in the theorem, namely that
R ⊆ G × M is an extent of the direct product if and only if the condition stated
in the theorem is fulfilled. The assumption that R is a dual bond is not needed.
However, with Proposition 76 this is easy, because obviously R is an extent if
and only if R = R∇∇ , and the latter is the case if and only if R(g) = R∇∇ (g)
is satisfied for all g ∈ G. But because according to Proposition 76 R∇∇ (g) =
I JJ
r
, this is exactly the condition in the theorem.
T
m∈gr I R(m )
Dense subcontexts
When the product R ◦ S was defined in Definition 62, we did not require that
the factors were bonds. Now we describe another situation in which this product
occurs naturally. Let K := (G, M, I) be a formal context and let G1 ⊆ G and M1 ⊆
M define a dense subcontext (G1 , M1 , I1 ), where I1 := I ∩ G1 × M1 , meaning that
3.4 Bonds and connections 159
Ah := hSI1 ∩ hT R .
Both hSI1 and hT R are extents of (G1 , M1 , I1 ) (the latter according to 1.), and
therefore Ah is the extent of some concept (Ah , AIh1 ) ∈ B(G1 , M1 , I1 ). When
(G1 , M1 , I1 ) is dense, then A0h = h0 , which is short for AIh1 = hS and AR
h = h .
T
{h} × hT ⊆ AIh1 S × AR
h,
which proves that T is contained in the right hand side of 3., i.e., in R ◦ S.
For the other inclusion let (X, Y ) ∈ B(G1 , M1 , I1 ) such that (h, n) ∈ Y S × X R .
Then h ∈ Y S and thus Y ⊆ hS = AIh1 , which implies that Ah ⊆ X and therefore
X R ⊆ AR h . This together with n ∈ X forces n ∈ Ah = h , i.e., (h, n) ∈ T .
R R T
To see that 3. is sufficient when 1. and 2. are given, we must infer from 3. that
each h ∈ G2 is reducible. So let Y := hS , X := Y I1 . Then Y is an intent of
(G1 , M1 , I1 ), h ∈ Y S and from 3. we get hT ⊇ X R . Let n ∈ hT \ X R . Then there
160 3 Parts, factors, and bonds
3.1
3.2
3.3
Kästner [210] shows how retracts of concept lattices (i.e., sublattices that are im-
age of an idempotent complete endomorphism) can be described by compatible
subcontexts in closed relations.
The dismantling of doubly irreducible elements was examined by Duffus and
Rival [108]. They also proved the uniqueness of the DI-kernel. A very simple
proof was given by Farley [123]. Berry and Sigayret [46] study complementary
contexts. Felde and Koyda [126] generalize the dismantling construction. Instead
of removing only doubly irreducible elements, they allow to remove intervals
[u, v], provided u is supremum-prime in (v] and v is infimum-prime in [u). They
prove that there is again a unique kernel.
Distributive lattices that are generated by their doubly irreducible elements
were examined by Monjardet & Wille [287].
It is possible to describe in terms of structure what happens if we add “a cross”
to the incidence relation I. Here, we shall limit ourselves to cardinality: The con-
cept lattice can become larger but it can also become smaller. A simple estimate
by Skorsky shows
1 |B(G, M, I ∪ {(g, m)})| 3
≤ ≤ .
2 |B(G, M, I)| 2
See also Kauer and Krupka [211]. The concept lattice can only become smaller
if g %
. m. Assume w.l.o.g. that (G, M, I) is clarified. If neither g % m nor g .
m, then, by Proposition 59, B(G, M, I) is a complete sublattice of B(G, M, I ∪
{(g, m)}).
3.4
Much of what we present in 3.4 has been moved there from Section 5.1 of the first
edition. The notion of a bond was introduced in [399] (under the French name
“liaison”). Relational Galois connections are from [141], but see also Xia [425],
Domenach and Leclerc [103]. The sublattices for the closed relations correspond-
ing to bonds (Proposition 65) were characterized by Kaarli et al. [206]. Kaarli
& Radecezki [207] explore the algebraic meaning of self-bonds.
162 3 Parts, factors, and bonds
Examples of irregular bonds occur in [141]. Krötzsch & Malik [238] remark
that both the ordinal scale M3 and the biordinal scale M2,1 have irregular self-
bonds (the concept lattices of which are called M3 and N5 in algebraic lattice
theory). These authors show connections to distributivity. Among other things,
they prove that Galois connections from and to completely distributive complete
lattices are regular and that a complete lattice which only has regular Galois
connections to other complete lattices must be distributive.
Chapter 4
Decompositions of concept lattices
A complex concept lattice can possibly be split into simpler parts. For this, the
mathematical modeling must prove its worth by providing effective and versatile
decomposition methods. Any such decomposition principle can be reversed to
create a method of construction. Therefore some of the following topics will be
taken up again in the next chapter with this second focus.
If a lattice can be represented as a sublattice of a direct product, this is called
a subdirect decomposition. The theory described in the previous chapter allows an
elegant description of these decompositions by means of the formal context. This
is the subject of the first section.
The tolerance relations introduced in 3.2 result in coverings of the concept
lattice by overlapping intervals. This fact will be used as a principle of decompo-
sition in the second section.
A surprisingly versatile context operation consists in inserting one context into
another one. We shall explain this in more detail in the third section and describe
the corresponding lattice construction, the substitution product. We then put some
effort into proving a decomposition theorem for this product. (Theorem 30).
In the fourth section we shall finally introduce the tensor product of complete
lattices by means of the direct product of the contexts. Similarly as in the case of
the direct product of lattices (which corresponds to the context sum), tensorial
decomposability is rare. Therefore, we transfer the idea of the subdirect prod-
uct, which we explain in 4.1, to contexts and obtain the notion of the subtensorial
decomposition of concept lattices.
The direct1 product of ordered sets has already been introduced in Definition 7.
We shall repeat it here for the special case of complete lattices:
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 163
B. Ganter, R Wille, Formal Concept Analysis, https://doi.org/10.1007/978-3-031-63422-2_5
164 4 Decompositions of concept lattices
Definition 65 Let T be an arbitrary index set. For a family (Lt , ≤t )t∈T of com-
plete lattices, the product is defined to be
× ×
(Lt , ≤t ) := Lt , ≤
t∈T t∈T
with
(xt )t∈T ≤ (yt )t∈T : ⇔ xt ≤t yt for all t ∈ T.
The lattices (Lt , ≤t ), t ∈ T , are the factors of the product, and the maps
πs : ×L
t∈T
t −→ Ls with πs ((xt )t∈T ) := xs
{ker πt | t ∈ T }
form a subdirect decomposition of (L, ≤). Conversely, for every subdirect decomposition
Θt , t ∈ T of (L, ≤), by
ι(v) := ([v]Θt )t∈T
an isomorphism of (L, ≤) onto a subdirect product of the factor lattices (L, ≤)/Θt , t ∈ T ,
is defined.
2For stylistic reasons, we will frequently leave out the adjective “complete”, i.e., in the following
“subdirect” should be replaced by “complete subdirect” where necessary.
4.1 Subdirect decompositions 165
(1,1)
d
(1,4) d @ d(6,5)
@
1
(5,4) d @ d (6,3) @ d (4,5) d
@ @
1
(2,4) d @ d (3,3) @ d (4,3) 5 d @ d6 d
@ @ @
(2,2) d 2 d 3 d @ d4 4 d @ d5
@ @ @ @
@ @
@ d (0,3) 2 d @d3
@ @ @ @
@ @
@d @d @d
@ @ @
(0,0) 0 0
Figure 4.1 The lattice on the left is a subdirect product of the two lattices on the right.
Proof The kernels of the canonical projections are congruences. Hence, in order
to prove the first part, we only have to show that their intersection is the trivial
congruence ∆. Two elements (vt )t∈T and (wt )t∈T of the direct product are differ-
ent if there is some s ∈ T with vs 6= ws , i.e., with πs (v) 6= πs (w), which
T is equiv-
alent to (v, w) 6∈ ker πs . Hence, from v 6= w it follows that (v, w) 6∈ t∈T ker πt ,
which was to be proved.
Now we show that the map ι defined in the second part has the properties
claimed. ι is a complete homomorphism, since for an arbitrary family (vj )j∈J of
elements of (L, ≤) we have
^ ^
ι vj = vj Θt
j∈J j∈J t∈T
^
= [vj ] Θt
j∈J t∈T
^
= ([vj ] Θt )t∈T
j∈J
^
= ι (vj ) .
j∈J
d d
@ @
d @d d @d
@ @ @ @
d @d @d d @d @d
@ @ @ @
d @d @d d @d @d
@ @
d @ d @
@ @ @ @
@ @d @ @d
@ @
@d @d
Figure 4.2 The two congruences represented here form a subdirect decomposition of the
lattice in Figure 4.1. The factor lattices by these congruences are precisely the factors of
the subdirect product represented in the figure above.
πs : ι((L, ≤)) −→ Ls , s ∈ T,
d
d @ @
@@
@
@
@
@
d @@ @d
@
@@ @d
@
@
@ @
@
@
@ @
@
d @@ @
@ @d
@
d @ @ @d
@ @d
@
@@ @
@ @
@
Q
Q
Q
Q
@
@
@d
@
@d
@
Figure 4.3 The nested line diagram helps to follow the definition of the subdirect product.
for the family C((L, ≤)) of all congruences of (L, ≤). Hence, these congruences
form a proper subdirect decomposition of (L, ≤).
The examination of subdirect decompositions can be carried out directly on
the context if we use the interplay between congruence relations, compatible and
arrow-closed subcontexts which we have developed in the preceding sections.
In order to do so, we must presuppose that the lattice (L, ≤) we examine is dou-
bly founded and thus isomorphic to the concept lattice of a reduced context K,
since in this case the congruences are in one-to-one correspondence to the arrow-
closed subcontexts of K.
Proof According to the observations preceding Theorem 13, t∈T ΘtS = ∆ holds
T
for a family Θt , t ∈ T of congruences if and only if t∈T Gt = G and t∈T Mt =
S
M holds for the corresponding arrow-closed subcontexts (Gt , Mt , I ∩Gt ×Mt ).
It is particularly easy to recognize the subcontexts belonging to subdirectly
irreducible factors of B(G, M, I). We must be aware that for every object g there
is always a smallest arrow-closed subcontext containing g. We shall call such a
subcontext a 1-generated arrow-closed subcontext. (The corresponding is true
for the attributes, but since in a reduced context every object is connected to an
attribute by a double arrow and vice versa, it suffices to concentrate on one of
the sets G or M , respectively).
Theorem 23 Every doubly founded complete lattice has a subdirect decomposition into
subdirectly irreducible factors.
Proof W.l.o.g. we may assume that (L, ≤) is the concept lattice of a reduced con-
text (G, M, I), and, since structural properties are preserved under isomorphism,
even assume that (L, ≤) = B(G, M, I). For g ∈ G let (Gg , Mg , I ∩ Gg × Mg )
denote the smallest arrow-closed subcontext of (G, M, I) containing g. Proposi-
tion 42 shows that this subcontext is reduced as well. According to Proposition
81 the corresponding concept lattice is subdirectly irreducible. Hence, together
with Proposition 80,
4.1 Subdirect decompositions 169
(Gg , Mg , I ∩ Gg × Mg ), g ∈ G,
d
fd
@ c
@d
b @ @ a b c d e f
d @d @ dd
1 × × %
. . × ×
5
a @ @ 2 .
% × × .
% ×
d @d @d
2 3 3 . .
% × × ×
e @ 4 × × × × %
. ×
d
1 @ 5 × × .
%
@ @
@ @d
@ 4
@d
Figure 4.4 Using the arrow relations in the context, we can examine which subdirect de-
compositions are possible for the concept lattice.
b a d c e f
KA AK A 6 6
A A A
A A AUA
AU AU ? ?
3 2 1 4 5
Figure 4.5 We recognize five 1-generated subcontexts: one for each object.
To close this section, we apply the theory we have developed to the example
given in the beginning (Figure 4.1). A representation of the lattice as a concept
lattice is presented in Figure 4.4. The arrow relations of the context are shown
in Figure 4.5 as a graph. We can read off that there are exactly five 1-generated
subcontexts. In the main, there is only one subdirect decomposition of this lattice
into subdirectly irreducible factors. The respective subcontexts are generated by
the objects 2, 4 and 5. We can add the subcontexts generated by the objects 1 or
170 4 Decompositions of concept lattices
d
@
d @d
@ @
d @d @d
@ @
d @d @d
@
d @
@ @
@ @d
@
@d
Figure 4.6 A coarser congruence may be chosen for the second congruence in the subdi-
rect decomposition from Figure 4.2.
It is, however, in the case of the tolerances not quite as easy as in the case of
topographic maps to explain the interconnection between the part maps. In the
general case, this is done through a family of adjoint pairs of mappings. It is
easier in the case of glued tolerances: here those maps ensue automatically.
Definition 68 Let Q be a complete lattice and, furthermore, let Lq := (Lq , ≤) be
a complete lattice for every element q ∈ Q. Let
(Lq | q ∈ Q)
with
x v y : ⇔ xmin ≤ ymin and ϕyxmin
min
x≤y
for all x, y ∈ Lq .
S
q∈Q ♦
Condition (4) says that for q ≤ r the two maps ϕrq and ψqr are mutually adjoint
W 11 on p. 19, as can easily
in the sense of Proposition be seen from Propositions
6 and 8. Hence, ϕrq is -preserving and ψqr is -preserving. If we denote the
V
boundary elements of (Lt , ≤) by 0t and 1t , respectively, we obtain in particular
ϕrq 0q = 0r and ψqr 1r = 1q .
Proposition 82
xmin ≤ ymin and ϕyxmin
min
x≤y
is equivalent to
xmax ≤ ymax and x ≤ ψxymax
max
y.
By (6) we obtain
y ≥ ϕqr x = ϕsr∨s x.
Hence, ϕqr x is an element of (Lr , ≤) ∩ (Lr∨s , ≤), and by (1) it follows that y ∈
(Lr∨s , ≤), i.e., s ≤ t. Since r ≥ xmin , we have s∧r ∈ [xmin , xmax ] and x ∈ (Ls∧r , ≤)
follows from (2). Together with (6) and (3) we obtain
ϕs∧r
q x = x.
Thus
r
ϕrs∧r x = ϕs∧r ϕs∧r r
q x = ϕq x ≤ y
Theorem 24 The sum of a Q-atlas is a complete lattice L := (L, ≤), in which infima
and suprema can be described as follows:
t u
_ ^
xt = ϕqrt xt and xt = ψsqt xt
t∈T t∈T
t∈T t∈T
with xt ∈ (Lqt , ≤), r := t∈T qt and s := t∈T qt . The complete lattices (Lq , ≤),
W V
q ∈ Q are precisely the blocks of a complete tolerance relation Θ of L, and q 7→ (Lq , ≤)
describes an isomorphism of Q onto L/Q; furthermore, we have
and
ψqr y = y u 1q for all y ∈ (Lr , ≤).
In this way we obtain a bijective assignment between the complete tolerance relations on
a complete lattice and the representations of this lattice as the sum of a Q-atlas.
we get ϕyxmax
max
x = y and, because of (4), x ≤ ψxymax max
y. Consequently, x v y holds
according to Proposition 82. Dually we infer that x = ψqr y always implies x v y.
We now show that the supremum has the form specified in the theorem: As-
sume that xt ∈ (Lqt , ≤) and xt v y for t ∈ T . By Proposition 82 andW(2) we
obtain y ∈ (Lymin ∨qt , ≤) for all t ∈ T and thus y ∈ (Lymin ∨r , ≤) for r := t∈T qt .
Therefore, from ϕxymin
min
xt ≤ y it follows that ϕqytmin ∨qt xt ≤ y because of (6) and
4.2 Atlas decompositions 173
min ∨r min ∨r
ϕqrt v ϕyqtmin ∨r xt = ϕyymin ymin ∨qt
∨qt ϕqt xt v ϕyymin ∨qt y = y
because of (5), (6) and (3). Thereby we have proved t∈T ϕqt xt v y. Since
r
W
The proved description of the suprema and infima immediately yields that Θ is
a complete tolerance relation of L. We can use Proposition 52 to show that the
(Lq , ≤) are precisely the blocks of Θ. This requires however to prove the maxi-
mality of the (Lq , ≤). Assume therefore that q ∈ Q and that y is an element with
B1 ≺ B2 in L/Θ implies B1 ∩ B2 6= ∅.
Let Σ(L) denote the smallest tolerance relation comprising all pairs (x, y) with
x ≺ y in L.
In the case of doubly founded lattices, a tolerance with overlapping neigh-
borhoods is called glued, and Σ(L) is called the skeleton tolerance. The factor
lattice L/Σ(L) is then called the skeleton of L. ♦
The intersection of any number of tolerance relations is again a tolerance re-
lation. Therefore Σ(L) is well-defined.
174 4 Decompositions of concept lattices
i.e., because of B1V≺ B2 , that [x]Σ(L) = B2 and thus x ≥ c. For this reason we
have that b ∨ c = {x | b < x ≤ d}, from which we can infer that b ≺ b ∨ c and
consequently (b, b ∨ c) ∈ Σ(L). Because of (b ∨ c)Σ(L) = c this yields c ≤ b ≤ d,
i.e., the desired contradiction b ∈ B2 .
It remains to be shown that Σ(L) is smallest among the tolerance relations
with overlapping neighborhoods. Hence, let Θ be any such tolerance relation
and x ≺ y a pair of neighboring elements of L. We must show that there is a
block of Θ containing x and y. This is certainly the case if x ∈ [y]Θ , i.e., we may
assume x 6∈ [y]Θ and in particular [x]Θ < [y]Θ . Now we consider an arbitrary
block [u, v] with [x]Θ ≤ [u, v] < [y]Θ . Because of u < yΘ ≤ y, y ∈ [u, v] would
immediately follow from y ≤ v. It would imply yΘu and thus a contradiction
u < yΘ . Hence, y 6≤ v and, because of x = (v ∧ y) Θ (v ∧ y Θ ) = v, we obtain xΘv
and thus x ∈ [u, v]. Hence, every block between [x]Θ and [y]Θ contains x. Since
the lower bounds of the blocks are closed under suprema, this also holds for
_
Bx := {B ∈ L/Θ | [x]Θ ≤ B < [y]Θ }.
Hence, this block must be a lower neighbor of [y]WΘ . Since Θ has overlapping
neighborhoods, it follows that Bx ∩ [y]Θ 6= ∅, i.e., Bx ∈ [y]Θ . If y 6∈ Bx , then
y ∧ Bx = x and thus x ∈ [y]Θ , contradicting [x]Θ < [y]Θ !
W
The corresponding block relation β(Σ(L)) can be described easily.
Σ := Σ(B(G, M, I))
be the skeleton tolerance. Then the following statements hold for the corresponding block
relation J := β(Σ):
a) J is the smallest block relation of (G, M, I) containing all pairs (g, m) with g %
. m.
b) J contains all pairs (g, m) with g . m or g % m.
Proof b) From g . m it follows that γg ∧ µm = (γg)∗ ≺ γg, i.e., (g, m) ∈ J
according to the definition of β. Dually we show that J furthermore contains all
pairs (g, m) with g % m.
a) Let K be a block relation comprising all pairs (g, m) with g %
. m and let fur-
thermore (A, B) and (C, D) be concepts with (A, B) ≺ (C, D). We want to show
4.2 Atlas decompositions 175
that (A, B) and (C, D) are related under the tolerance relation β −1 (K) belonging
to K. For this purpose, we consider an object g ∈ C and an attribute m ∈ B with
(g, m) ∈/ I. Since the context is doubly founded, we find an object h with h0 ⊇ g 0
and h . m, i.e., in particular (h, m) ∈/ I. We again make use of the fact that the
context is doubly founded to get an attribute n with n0 ⊇ m0 and h % n and thus
h%. n (since µn ≥ µm ≥ (γh)∗ ). Consequently, (h, n) ∈ K, which is equiva-
lent to (γh, γh ∧ µn) ∈ β −1 (K), or shorter (γh, (γh)∗ ) ∈ β −1 (K). From that we
infer ((A, B) ∨ γh, (A, B) ∨ (γh)∗ ) ∈ β −1 (K), i.e., ((C, D), (A, B)) ∈ β −1 (K), as
claimed.
This is compatible with Definition 68. The maps postulated there ensue canon-
ically in the glued case, as the following proposition shows.
for q = q0 ≺ q1 ≺ · · · ≺ qm = r.
176 4 Decompositions of concept lattices
{q ∈ Q | x ∈ (Lq , ≤)}
1st case: q1 ∨ r1 = qm .
Since the smallest elements of (Lq1 , ≤) and (Lr1 , ≤) belong to (Lq0 , ≤), the supre-
mum 0q1 ∨ 0r1 can be formed in (Lq0 , ≤). This element belongs to (Lq1 , ≤) as well
as to (Lr1 , ≤); i.e., the two lattices are not disjoint. Because of 4) we have
i.e., (Lq0 , ≤) and (Lqm , ≤) are not disjoint either, and 0qm has to be an element
of (Lq0 , ≤). Because of 5), 0qm therefore belongs to all lattices (Lqi , ≤), i ∈
{0, . . . , m} and to all lattices (Lrj , ≤), j ∈ {0, . . . , n}. The sets (Lq0 , ≤) ∩ (Lqi , ≤)
and (Lq0 , ≤) ∩ (Lrj , ≤) are therefore all nonempty. Since by 1) they are order
ideals, the elements 0qi and 0rj , respectively, must all belong to (Lq0 , ≤).
Hence, for x ∈ (Lq0 , ≤) it follows that
(ϕqqm−1
m
◦ · · · ◦ ϕqq10 )x = (. . . ((x ∨ 0q1 ) ∨ 0q2 ) ∨ . . .) ∨ 0qm ,
and all those suprema are being formed in (Lq0 , ≤). Therefore,
ϕqqm
0
x = x ∨ 0q1 ∨ · · · ∨ 0qm = x ∨ 0qm
and correspondingly
q 1 = s 1 ≺ s 2 ≺ · · · ≺ s j = q 1 ∨ r1
r 1 = t 1 ≺ t2 ≺ · · · ≺ tk = q 1 ∨ r 1
ϕqqm
m−1
◦ · · · ◦ ϕqq12 = ϕssl−1
l
◦ · · · ◦ ϕss12
and
t
ϕrrn−1
n
◦ · · · ◦ ϕrr21 = ϕtl−j
l−j−1
◦ · · · ◦ ϕtt12 .
Now, by insertion we obtain
ϕqqm
m−1
◦ · · · ◦ ϕqq10 = ϕrrnn−1 ◦ · · · ◦ ϕrr10 .
Thereby we have shown that the definition of ϕrq is independent of the choice of
the chain.
Furthermore, we have to prove (4), i.e., that the pairs of maps ϕqr , ψqr , q ≤ r
are adjunctions. If q ≺ r, then this follows immediately from the definition, since
for x ∈ (Lq , ≤), y ∈ (Lr , ≤) we have
ϕrq x ≤ y ⇔ x ∨ 0r ≤ y ⇔ x ≤ y
and dually
x ≤ ψqr y ⇔ x ≤ y ∧ 1q ⇔ x ≤ y.
Thus, we get
ϕqqm−1
m
◦ · · · ◦ ϕqq10 x ≤ y
⇔ ϕqqm−1
m−2
◦ · · · ◦ ϕqq10 x ≤ ψqqm−1
m
y
.
⇔ ..
⇔ x ≤ ψqq01 ◦ · · · ◦ ψqqm−1
m
y.
Condition (5) follows immediately from parts (1) and (3) of the definition. What
remains to be shown is condition 6) of Definition 68. In order to do so, we first
consider the case q ≺ r, for which from x ∈ (Lq , ≤) ∩ (Lq∨s , ≤) it follows that:
r∨s
ϕrq x = x ∨ 0r = x ∨ 0r ∨ 0q∨s = x ∨ 0r∨s = ϕq∨s x
(if q ∨ s ≺ r ∨ s does not hold, then the last of these equalities is inferred as
above for ϕqqm
0
). The general case is obtained by concatenation along a chain of
neighbors.
178 4 Decompositions of concept lattices
≤ being the transitive closure of the union of the orders on the summands. ♦
Theorem 27 The sum of a Q-atlas with overlapping neighbor maps is a complete lattice
L where the summands (Lq , ≤), q ∈ Q are precisely the blocks of a complete tolerance
relation Θ and where q 7→ (Lq , ≤) describes an isomorphism of Q onto L/Q.
Conversely, in a complete lattice L the blocks of a tolerance Θ with overlapping neigh-
borhoods, for which Q := L/Θ is of finite length, always form a Q-atlas with overlapping
neighbor maps whose sum is L.
Proof First of all, we shall prove that the order v of the Q-atlas, which is de-
scribed by Proposition 83, is equal to the transitive closure ≤ of the union of
the orders on the summands. According to the definition, v on the summands
(Lq , ≤) coincides with their respective orders, which is the reason why from
x ≤ y always follows x v y. If, conversely, x v y, i.e., xmin ≤ ymin and ϕyxmin
min
x ≤ y,
then for
xmin = q0 ≺ q1 ≺ · · · ≺ qm = ymin
as in Proposition 83 it follows that
ϕxymin
min
x = (. . . ((x ∨ 0q1 ) ∨ 0q2 ) ∨ . . .) ∨ 0qm ,
Therefore, Proposition 83 yields the assertions of the first part of Theorem 24.
Furthermore, we get that in a complete lattice L the blocks of a glued tolerance
Θ form a Q-atlas. It only remains to be shown that this Q-atlas is in fact a Q-atlas
with overlapping neighbor maps. Conditions 0), 1), 2) and 3) of Definition 70
are obviously satisfied. From 0q ≤ x ≤ 1q and 0r ≤ x ≤ 1r it follows that
0q∨r = 0q ∨ 0r ≤ x ≤ 1q ∧ 1r = 1q∧r ,
which proves 4). 5) can be seen from the fact that q ≤ r ≤ s and 0s ≤ x ≤ 1q ,
because of 0r ≤ 0s and 1q ≤ 1r , immediately yield 0r ≤ x ≤ 1r .
Theorem 27 can be applied quite practically for the representation in dia-
grams, provided that the lattice which is to be represented has a tolerance with
overlapping neighborhoods. This can be checked by entering the arrow relations
into the context and enriching the relation
4.2 Atlas decompositions 179
J := I ∪ . ∪ %
in accordance with the conditions in Definition 59, until a block relation is ob-
tained (namely β(Σ)). According to Corollary 1, we obtain the blocks (“maps”)
as concept lattices of subcontexts. Diagrams are created from these. The overlaps
can be read off, even in the case of a reduced labelling of the individual maps,
since every concept is stated with its correct extent and intent.
According to Theorem 27, the lattice is uniquely described by this set of dia-
grams. The correctness of the atlas can be verified by means of the conditions in
Definition 70.
We shall demonstrate this using the example of a lattice of subgroups. Fig-
ure 4.7 shows a computer-generated diagram, which was taken from a book on
orthomodular lattices [205].
19
18 17 16
14 15 13 11 12
9 10 8 6 7
5 4 3 2
10 14 15 16 17 18
2 .
% × × × × ×
3 × .
% × × %
. ×
4 × .
% × .
% × ×
5 × × × .
% .
% ×
6 × .
% × × ×
11 . × × .
%
Figure 4.8 The standard context for the lattice from Figure 4.7
The standard context (cf. Page 32) for this lattice, including the arrow rela-
tions, is presented in Figure 4.8. A short examination shows that J := I ∪ . ∪ %
is already a block relation, i.e., that it is equal to β(Σ).
The concept lattice of (G, M, J) is a three-element chain with formal concepts
The subcontexts belonging to the blocks of the relation are represented to-
gether with their concept lattices in Figure 4.9. In the case of the lattice presented
in the middle of this figure, it can be easily seen how the blocks overlap. The
smallest element of this block is the concept with the extent {2}. We discover it
in the lower lattice on the right side. The largest element of the lower lattice has
the intent {15, 18}. It can easily be found in the middle lattice.
Hence, the lower and the middle lattice have the five elements of the interval
[({2}, {14, 15, 16, 17, 18}), ({2, 3, 4, 5}, {15, 18})]
in common. Analogous are the middle and the upper lattice, which overlap in
the interval
[γ6, µ18],
also having five elements.
In the present case, it proves to be particularly convenient that we have chosen
congruent diagrams for the overlap areas. This makes it possible to superimpose
the individual part-diagrams (Figure 4.10) and thus to obtain a new diagram for
the subgroup lattice (Figure 4.11), which reflects the structure particularly well,
due to its construction method.
4.2 Atlas decompositions 181
d
@
@
14 16 17 18 18 d 17 d @ 16
@d
2 × × × × A@ @
3 × × A@ @
4 × ×
d A d14@ d
A @ @
5 × × @d
6 × × × × 4 @5 A 3 11
11 × × @ A
@A
@A d
2,6
18
d
A@
A@
15 d 17 d A d14@ d16
A @
14 15 16 17 18
2 × × × × × A@ @ A
3 × × × A@ @ A
4 × × × A @ @A
5 × × × d Ad @ d @A d
6 × × × × 4 @5 A 3 6
@ A
@A
@A d
2
15,18
d
A@
A@
10 d 17 d A d14@ d16
A @
10 14 15 16 17 18
A@ @ A
2 × × × × × A@ @ A
3 × × × × A @ @A
4 × × × × d Ad @ d @A d
5 × × × × 4 @5 A 3 2
@ A
@A
@A d
e
@
@
@
18 17 @ e16
` e e @
`
``
A@ @
` ` A @ @
`
18 ` `
A @ @
e 4 e 5 A e14@
` `
@e
`
@e
@
` ` ` ` ` ` 3 11
A @` ` ` ` ` ` @` ` ` A
A@ @ A
` A` @` ` ` ` ` @ A
15 e 17 e` A e` @ @e @ @A e
` ` 14 16 `
` ` A@ @ A
` ` ` 2,6
`` A@ ``
@ A
` `
15,18 ` ` `
A @ @ A
e 4 e 5A e @ @e @ @A e`
` ` ` 3
A@
` `` @ ` ` ` A` ` ` 6
A ``
@
` ` @` A
`
` ` A` @` ` ` ` ` ` @ A
10 e 17 e A e @ e @ @ @A e
14 16
` `` 2
`
A@ @ A
A@ ``
@ A
A @ @A ` `
e Ae @ @e @@A e`
4 @ 5 A 3 2
@ A
@A
@A e
@
Figure 4.10 Atlas of the part diagrams. The dotted lines link equal concepts in the different
part diagrams. If the diagram is contracted along those lines, we obtain the diagram in Figure
4.11.
4.2 Atlas decompositions 183
f
@
@
@
@
18 f 17 f @@ f16
A@ @
A@ @
A @ @
A @ @
15 f f A f14@@f @@f
@
A @ A 11
A@ @ A
A @ @ A
A @ @A
10 f f A f @
@ f @A f
@
A@ @ A 6
A@ @ A
A @ @ A
A @ @A
f A f @
@ f @A f
@
4 @ 5 A 3 2
@ A
@ A
@A
@A f
@
Figure 4.11 The same lattice as in Figure 4.7, but with a diagram which better reflects the
structure.
184 4 Decompositions of concept lattices
4.3 Substitution
In the case of the substitution sum a context is inserted into another one at the
place of an “empty cell”, i.e., a non-incident object-attribute pair. We can visual-
ize the construction by imagining the respective row and column suitably mul-
tiplied, so that there is room for the context which is to be inserted. For reasons
of convenience we presuppose that the two contexts are disjoint and non-empty,
and thus we obtain the following definition.
Definition 72 Let K1 = (G1 , M1 , I1 ) and K2 = (G2 , M2 , I2 ) be contexts and
/ I1 be a non-incident object-attribute pair in K1 . We presuppose that
(g, m) ∈
6 M2 and (G1 \ {g}) ∩ G2 = ∅ = (M1 \ {m}) ∩ M2 . We define substitu-
G2 6= ∅ =
tion sum of K1 with K2 on (g, m) to be the context
K1 (g, m) K2 := (G, M, I)
I := {(h, n) ∈ I1 | h 6= g, n 6= m} ∪ G2 × g I1 ∪ mI1 × M2 ∪ I2 .
M2
m
g G2 K2
K2 -
K1
Figure 4.12 To form the substitution sum K1 (g, m) K2 , the context K2 is inserted into “the
empty cell” (g, m) of K1 . The hatchings in the resulting context are meant to indicate that
every object 6∈ G2 is incident either with all or with no element from M2 and, dually, that
all objects from G2 have the same intents with regard to the attributes 6∈ M2 .
From the definition it immediately follows that the substitution sum is restrict-
edly associative:
4The cases K2 ∼ = ({g}, {m}, ∅), when K1 (g, m) K2 = K1 , as well as K1 = ({g}, {m}, ∅) are
admitted. In the following we shall consider proper substitution sums only.
4.3 Substitution 185
Proposition 84
K1 ×, K1 ∅ and
K1 ×
× K2 ∅ K2 ∅ K2
We shall examine how the concept lattice of the substitution sum is related to
those of the summands. It turns out that the concept lattice of the second sum-
mand is “hung up” several times in the concept lattice of the first summand,
similarly to the sails of a ship in the rigging. (see Figure 4.13).
Rigging
Sails aa
J a
@ J aa
@ Ja J
aa J
aa Ja
b aa
J
J a
Ja J
aa J
aa J
(G, M, I) = K1 (g, m) K2 ,
defined as
U := B(G, M, I \ I2 ),
is a complete sublattice of B(K1 (g, m) K2 ) that is isomorphic to B(K1 ). U contains in
particular all concepts which are ≥ a or ≤ b.
Proof Proposition 57 shows that I \ I2 is closed. The context (G, M, I \ I2 ) is up
to clarification identical to K1 , i.e., the isomorphy follows from Theorem 17. A
concept (A, B) ≤ b satisfies A ⊆ mI1 and therefore A × B ∩ I2 = ∅, which implies
(A, B) ∈ U .
The remaining concepts of K := (G, M, I), i.e., those
which do not belong to B(G, M, I \ I2 ), all contain “one
cross from I2 ” and thus are entirely contained in the M2 g 0
subcontext (G2 ∪mI1 , M2 ∪g I1 ). This subcontext, accord-
ing to the definition of K, is the sum of K2 and K3 := ∅
G2 K2
@
@
(m , g , I3 := I ∩ (m × g )), i.e., the concept lat-
I1 I1 I1 I1 @
m0 @ K3
tice of this subcontext is isomorphic to B(K2 ) × B(K3 ). @
@
We do not claim that we thereby obtain a sublattice,
∅
but by Proposition 38 we know that we find an order-
embedding of this concept lattice into B(K) by assign-
ing the concept (AII , AI ) of K to every concept (A, B)
of K2 + K3 . The proposition just mentioned suggests a
further order-embedding, namely (A, B) 7→ (B I , B II ). In the present case, how-
ever, this yields the same map, since AI = B or B I = A for every concept (A, B)
of K2 + K3 .
We denote the image of this mapping by P and summarize:
Proposition 86 The map
is an order-embedding, mapping the concept with the extent G3 (= mI1 ) onto b and the
concept with the intent M3 (= g I1 ) onto a. The range P covers all concepts which do
not belong to U .
We call P the sails of the substitution sum. The following theorem shows that
the two parts, rigging and sails, determine the structure of the concept lattice of
a substitution sum. However, we must indicate how U and P are joined together.
Before doing so, we shall introduce the corresponding lattice construction.
Definition 73 For complete lattices U and W , |W | > 1 and elements a, b ∈ U
with a 6≤ b, we define the substitution product U (a, b) W of U and W on (a, b)
to be the concept lattice of the (proper) substitution sum
4.3 Substitution 187
and
g ≤ m in U, if g ∈ U, m ∈ U
g ≤ b in U, if g
∈ U, m ∈ W
gIm ⇔
a ≤ m in U,
if g ∈ W, m ∈ U
g ≤ m in W, if g ∈ W, m ∈ W
c c
cHH cHH
c BZ HH c BZ HH
c Z
Bc Z c
Hc c Z Hc
C c c c
Bc Z c
H bJ bJb HH A C
cH C b HA cH C
HH AA C J b bb C
b c c c c
Jb c c
Jb c c c ca
H a b J JJ J b b b H
C H cH C H cHb
H H bb J J
bbJb
C AA cHH c c C AA cHHb c b
b Jb
J bJcJb
JcJ c
Cc Z c Cc Z c
HH Z B c c c@c @ HH Z B c
HHZB c HHZB c
U H c W@c
@
U (a, b)W H c
B(K1 (g, m) K2 ) ∼
= B(K1 ) (γg, µm) B(K2 ).
188 4 Decompositions of concept lattices
p = (p ∧ a) ∨ (p ∧ b) as well as p = (p ∨ a) ∧ (p ∨ b)
for all p ∈ P ,
(Subst 4) If u ∈ U and p ∈ P \ U , then u ≤ p ⇒ u ≤ b and u ≥ p ⇒ u ≥ a.
Proof In Propositions 85 and 86 we have already shown that B(K1 (g, m)K2 )
has the properties (Subst 1) and (Subst 2), and furthermore that P is order-
isomorphic to the direct product of B(K2 ) and B(K3 ). This yields the first part
of (Subst 3). A concept p = (X, Y ) of K1 (g, m)K2 which does not belong to U
satisfies X ⊆ G2 ∪ mI1 and Y ∩ M2 6= ∅. Every subconcept of it which belongs to
U must therefore have an extent which is entirely contained in mI1 , i.e., it is ≤ b.
This proves (Subst 4). Now we can infer the second part of (Subst 3), since an
upper bound of (p ∧ a) ∨ (p ∧ b) which is less than or equal to p must be contained
in P and must therefore be equal to p.
We now assume, conversely, that L has the properties (Subst 1) – (Subst 4).
First we derive some information from the structure of P . From (Subst 3) we
infer that P ⊆ [a ∧ b, a ∨ b]. The maps p 7→ p ∨ b and q 7→ q ∧ a are isomorphisms
between P ∩ (a] and P ∩ [b) which are inverse to each other, since it holds for
p ≤ a that
(p ∨ b) ∧ a = (p ∨ b) ∧ (p ∨ a) = p,
and dually. In this way we do not only obtain the isomorphism σ : W → P ∩ (a]
postulated in (Subst 3) but a further isomorphism τ : W → P ∩ [b) by virtue of
τ (x) := σ(x) ∨ b. From (Subst 1) we can see that the elements of P ∩ (a], with
the exception of the boundary elements a and a ∧ b, do not belong to U . The
corresponding is true for P ∩ [b).
In order to prove the isomorphy of L with U (a, b) W , we use the Basic The-
orem on Concept Lattices. The context defining U (a, b) W has the object set
G = (U \ {a}) ∪ (W \ {0W }) and the attribute set M = (U \ {b}) ∪ (W \ {1W })
(cf. the explanation following Definition 73). The maps
γ̃ : G → L, µ̃ : M → L,
if x ∈ U if x ∈ U
x x
γ̃(x) := and µ̃(x) :=
σ(x) if x ∈ W τ (x) if x ∈ W ,
B(K1 (g, m) K2 )
of a proper substitution sum has (using the notations in Theorem 28) the following prop-
erties:
(Subst 5) The sails P are isomorphic to a direct product
P ∼
= (P ∩ (a]) × (P ∩ (b]).
x≤y ⇔ x≤y∧a ⇔ x ∨ b ≤ y.
(Subst 7) Each element of P ∩ (a] is the supremum of object concepts in P ∩ (a], and
each element of P ∩ W[b) is the infimum of attribute concepts in P ∩ [b).V
(Subst 8) If 1W is -irreducible, then a is an object concept. If 0W is -irreducible,
then b is an attribute concept.
190 4 Decompositions of concept lattices
B(K) ∼
= B(K1 ) (γg, µm) B(K2 ).
ψγg, if g ∈ GU , falls m ∈ MU ,
ψµm,
γ1 (g) := and µ1 (m) :=
a, if g = ga , b, if m = mb ,
by the Basic Theorem are sufficient for the isomorphy of B(K1 ) with U , since
evidently g I m ⇔ γ1 g ≤ µ1 m.
The context K2 := (G2 , M2 , I2 ) is explained through
G2 := G \ GU , M2 := M \ MU , I2 := I ∩ G2 × M2 .
γ2 g := ψγg, µ2 m := (ψµm) ∧ a.
Proof If
ψ : B(K) → U (a, b) W
is an isomorphism, so that ψ −1 satisfies the extra condition, then we define sub-
context K0 := (G0 , M0 , I ∩ G0 × M0 ) through
G0 := {g | ψγg ∈ U ∪ (a]}
M0 := {m | ψµm ∈ U ∪ [b)}.
192 4 Decompositions of concept lattices
We only have to prove that K0 is a dense subcontext, since in this case Proposi-
tion 88 yields the rest of the assertion.
For this purpose we show that γG0 is -dense. The corresponding asser-
W
tion M0 isWproved dually. Since ψ is an isomorphism,
W we can prove instead that
ψγG0 is -dense in U (a, b) W . ψγG is certainly -dense, i.e., every element
s ∈ U (a, b) W can be represented as a supremum
_
s= X, X := (s] ∩ ψγG.
M2
G2 K2
K∗1
K = K1 ( , ) K2 .
L := L1 (c1 , d1 ) L3 ∼
= L2 (c2 , d2 ) L4
and that K is the (reduced) standard context for L. We can apply Theorem 29,
since the extra condition is irrelevant because of the additional preconditions.
Hence, K is in two ways a substitution sum:
K∼
= K1 (g1 , m1 ) K3 ∼
= K2 (g2 , m2 ) K4 ,
with
B(Ki ) ∼
= Li and Ki =: (Gi , Mi , Ii )
for i ∈ {1, . . . , 4}. The presuppositions of the proposition guarantee that K3 and
K4 have neither full nor empty rows or columns.
Hence, the isomorphy of the substitution products corre- W
sponds to the isomorphy of two substitution sums. In the fol- L1 1 L2
lowing, we shall make use of this circumstance. There are, how-
@
R
@
W2 W3
ever, some complications, since the result for substitution sums
which corresponds to Proposition 89 does not hold in general. L@@
R
4 L
W4 3
If, however, we manage to find contexts L1 , L2 , L3 , L4 with
K1 = L1 ( , ) L2 , K 2 = L1 ( , ) L3 , L1
K1 K2
K3 = L3 ( , ) L4 , K 4 = L2 ( , ) L4 , @
R
@
L2 L3
then the statement of the proposition results from Theorem 28.
K@
R
@
4 K3
This is true for some special cases. These will be dealt with first. L4
M4
K3
G4
K∗0
K∗2 M1∗
G∗1
K∗1 = K2 ( , ) K∗0 , K1 = K2 ( , ) K0 , K4 = K0 ( , ) K3
If, as a fourth context, we add the trivial context 2 := ({g}, {m}, ∅), we obtain
the desired refinement with
K1 = K2 ( , ) K0 K2 = K2 ( , ) 2
K3 = 2 ( , ) K3 K4 = K0 ( , ) K3
(cf. the left diagram of Figure 4.17). Of course, thereby we have also dealt with
the converse case G4 ⊂ G3 , M4 ⊂ M3 .
K3
K2 K0
K1 K2 K1 K2
K∗∗
@
2
R
@ @
R
@
K0 0 K4 K3
M1∗
K@ K@
2
R
@
4 K3 R
@
4 K3
K3 K4
G∗1
Case 2: G3 ∩ G4 = ∅ = M3 ∩ M4 .
We define (middle figure in 4.17)
G0∗∗ := G \ (G3 ∪ G4 ),
M0∗∗ := M \ (M3 ∪ M4 ).
Let K0 be the subcontext with the object set G0 := G0∗∗ ∪ {g, h} and the attribute
set M0 := M0∗∗ ∪{m, n}, with (g, m) and (h, n) being non-incident object-attribute
pairs of K3 or K4 , respectively.
We recognize that K0 (g, m) K4 is equal to K1 . Furthermore, K2 = K0 (h, n) K3 .
Once again by using the trivial context 2 we obtain the refinement represented
in the diagram on the right, which yields the statement of the proposition.
The cases G3 ⊆ G4 , M4 ⊆ M3 (resp. dually) and G3 ∩ G4 6= ∅, M3 ∩ M4 = ∅
turn out to be trivial: If, for instance, G3 ⊆ G4 and m ∈ M3 \M4 , then m0 ∩G4 = ∅
or m0 ∩ G4 = G4 and consequently m0 ∩ G3 = ∅ or m0 ∩ G3 = G3 , contrary to our
presuppositions. If M3 ∩ M4 = ∅, then G3 ⊆ G4 is obviously impossible, since
otherwise K3 would have constant columns. So, if g ∈ G3 ∩ G4 , h ∈ G3 \ G4 and
m ∈ M4 , then from g I m it immediately follows that h I m and thus h I n for all
n ∈ M4 , which in turn necessitates g I n for all n ∈ M4 . Similarly, from (g, m) ∈
/I
it follows that g 0 ∩ M4 = ∅. This means that K4 would contain a full row or an
empty row, which is contrary to the presuppositions.
What remains is the case that the two subcontexts K3 and K4 intersect non-
trivially. We can proceed similarly as we have done so far and introduce contexts
L1 , . . . , L4 , as presented in Figure 4.18:
K3 L∗3
L1
L4 K1 K2
@
R
@
K4 L∗2 L2 L3
L∗1 K@
R
@
4 K3
L4
being chosen arbitrarily. The new object and the new attribute are reducible in
K as well as in the subcontexts K1 , . . . , K4 , i.e., the respective concept lattices are
isomorphic. If by L+ 4 we denote the context resulting from L4 , we obtain
K 1 = L1 ( , ) L2 K2 = L1 ( , ) L3
K3 = L3 ( , ) L+
4 K4 = L2 ( , ) L+
4,
L∼
= ((. . . (N1 ( , ) N2 ) . . .) ( , ) Nn−1 ) ( , ) Nn
be two decompositions of L into indecomposable factors M1 , . . . , Mm or N1 , . . .,
Nn , respectively (the names of the elements in brackets are irrelevant for the
proof). Assume that n is the largest possible length which this kind of decom-
position of L can have. We proceed by induction on n.
According to Proposition 89 there are lattices W1 , . . . , W4 with
Mm ∼
= W3 ( , ) W4 and Nn ∼
= W2 ( , ) W4 .
Since Mm and Nn are substitutionally indecomposable, |W2 | = |W3 | = 2 or
∼ W4 ∼
|W4 | = 2. In the first case Mm = = Nn and
∼ W1 ∼
(. . . (M1 ( , ) M2 ) . . .)Mm−1 = = (. . . (N1 ( , ) N2 ) . . .)Nn−1 ,
decomposition of L with more than n factors. By induction we can infer that all
decompositions of W1 into indecomposable factors have the same number k of
factors. This number, however, must equal n − 2, since
∼ (. . . ((N1 ( , ) N2 ) ( , ) N3 ) . . .) ( , ) Nn−1 .
W 1 ( , ) Mm =
Also by the induction hypothesis, every decomposition of this lattice has pre-
cisely n − 1 factors.
×K
t∈T
t := ( × G , × M , ∇),
t∈T
t
t∈T
t
with
g∇m : ⇔ ∃t∈T gt It mt
for g := (gt )t∈T and m := (mt )t∈T . ♦
We had introduced this definition already in Section 1.4 for the special case of
two factors:
K1 × K2 := (G1 × G2 , M1 × M2 , ∇),
(g1 , g2 )∇(m1 , m2 ) : ⇔ g1 I1 m1 or g2 I2 m2 .
For reasons of simplicity we will use the following abbreviations throughout this
section:
G := ×G ,
t∈T
t M := ×M ,
t∈T
t g := (gt )t∈T and m := (mt )t∈T .
A tiresome complication in the notation stems from the trivial case of the “full
rows” and “full columns”. We use the notation introduced in Section 3.3
:= M ∇ × M ∪ G × G∇ .
However, the reader can assume without great loss of generality that full rows
and columns do not occur, and can set G∇ = M ∇ = = ∅ everywhere.
Every trivial context having only one concept, i.e., every context of the form
(G, M, G×M ), acts like a zero element for the direct product: If one of the factors
is trivial, so is the product. Occasionally, we have to exclude this case.
c d e
a b
3 ×
1 ×
2 × 4
5
×
×
=
K1
K2
a a a b b b a b a b a b
c d e c d e c c d d e e
1 3 × × × × 1 3 × × × ×
1 4 × × × × 2 3 × ×
1
2
5
3
×
×
× ×
×
×
= 1 4 ×
2 4
×
×
× ×
×
2 4 × × 1 5 × × × ×
2 5 × × 2 5 × ×
K1 × K2 K1 × K2
e
@
@
@
(a, c) ee(a, d) @
@ e (a, e)
@@ @
@ @ @
@ @ @
(b, c) (b, d) @@ e(b, e)
e @@ e @
@ e
(2,3) @ @ (2,4) (2,5)
@ @
@ @
@e
(1,3)@ e (1,4) @
@ e (1,5)
@
@
@
@@e
B(K1 × K2 )
Figure 4.20 The concept lattices of the direct product of the contexts from Figure 4.19.
4.4 Tensorial decompositions 199
ε1 (1) = ε2 (1)
c
@
@
1 c c c @@c
1
c
@ @ @
@ @ @
c ε2 (x) c @c @ cε2 (y)@
@ c ε2 (z) c cy@ c z
@
w ε1 (w)
@ @ x
c
@ @ @ @
@ @
0
0 c @c c @c
@ @
@
@
@@c
ε1 (0) = ε2 (0)
∇t := {(g, m) ∈ G × M | gt It mt } ∪
εt : B(Kt ) → B(G, M, ∇)
with
εt (A, B) := ({g ∈ G | gt ∈ A} ∪ M ∇ , {m ∈ M | mt ∈ B} ∪ G∇ )
is a canonical lattice-embedding.
Proof The fact that ∇t is closed can be proved easily, for instance by means of
Proposition 57: If (g, m) ∈ ∇ \ ∇t , then in particular m∇ 6= G, i.e., we can choose
an object g̃ 6∈ m∇ . Using this object, we define an object h to be
g̃s if s 6= t
hs := .
gs if s = t
h ∈ M ∇ or
g ∇ ⊆ h∇ ⇔ .
gtIt ⊆ hIt t for all t ∈ T
200 4 Decompositions of concept lattices
Proof “⇒”: We presuppose the negation of the right side. Assume that m 6∈ h∇
and nt ∈ gtIt \ hIt t for some t ∈ T . Consider the attribute m̃, defined by
ms if s 6= t
m̃s := .
nt if s = t
Proposition 92
g . m ⇔ ∀t∈T gt . mt ,
g % m ⇔ ∀t∈T gt % mt .
Proof We prove only the first statement. First of all we notice that, because of
(g, m) 6∈ ∇ ⇔ ∀t (gt , mt ) 6∈ It , we can limit ourselves to non-incident pairs. If,
for some t, gt . mt does not hold, there must be an object ht ∈ Gt with gt0 ⊆ h0t ,
gt0 6= ht0 and (ht , mt ) 6∈ It . The object g̃, defined by
gs if s 6= t
g̃s :=
ht if s = t
satisfies
g ∇ ⊆ g̃ ∇ , g ∇ 6= g̃ ∇ and (g̃, m) 6∈ ∇,
which yields ¬(g . m).
Analogously, we infer the converse direction: If gt . mt holds for all t ∈ T ,
then we certainly have (g, m) 6∈ ∇ and we only have to consider an object h with
g ∇ ⊆ h∇ , g ∇ 6= h∇ . By Proposition 91 we obtain gs0 ⊂ h0s , gs0 6= hs0 for some s ∈ T ,
from which, because of gs . ms , it immediately follows that hs Is m and thus
h∇m.
Together with Proposition 15 this yields
Corollary 3 An object g of a direct product is irreducible if and only if all gt are irre-
ducible. The corresponding is true for attributes.
The direct product of reduced contexts is reduced, the direct product of doubly founded
contexts is doubly founded.
Our interest lies in the concept lattice of the direct product. We shall call this
lattice the tensor product of the factor lattices B(Kt ). Thereby we obtain a new
lattice construction and thus a new decomposition principle. However, in order
to do so we have to show that the tensor product is independent (up to iso-
morphism) of the choice of the underlying contexts Kt . This is the result of the
theorem which follows the next definition.
Definition 75 The tensor product of complete lattices Lt , t ∈ T , is defined as
×
O
Lt := B( (Lt , Lt , ≤)),
t∈T t∈T
4.4 Tensorial decompositions 201
L1 ⊗ L2 := B(L1 × L2 , L1 × L2 , ∇)
with
(g1 , g2 )∇(m1 , m2 ) : ⇔ g1 ≤ m1 or g2 ≤ m2 . ♦
f
@
f f f f @f
@
@ @ @ @
f ⊗ f f @f f @f @f @ @f
@ ∼
@ @
=
@ @ @
f @f @f f @f
@ @ @
@
@@f
Theorem 31 The concept lattice of a direct product of contexts is isomorphic to the tensor
product of the concept lattices of the factor contexts:
× Kt ) ∼
O
B( = B(Kt ).
t∈T t∈T
For the proof we use the Basic Theorem on Concept Lattices. According to Def-
inition 75, the tensor product t∈T B(Kt ) is the concept lattice of the context
N
(G, M, ∇) with
G=M = B(Kt )×t∈T
and
(At , Bt )t∈T ∇ (Ct , Dt )t∈T ⇔ ∃t∈T At ⊆ Ct .
In order to prove the isomorphy we claimed, we have to give maps
γ̃ : G → B( ×K )
t∈T
t and µ̃ : M → B( ×K )
t∈T
t
{g | ∀t gt ∈ At }∇ = {m | ∃t mt ∈ At0 } and
∇
M ∪ {g | ∀t gt ∈ At } = {m | ∃t mt ∈ At0 }∇ .
The inclusions ⊆ are trivial. Therefore, let m be an attribute with mt 6∈ A0t for
all t ∈ T . Then for every t ∈ T there exists an object gt ∈ At with (gt , mt ) 6∈ It ;
and thus g := (gt )t∈T satisfies (g, m) 6∈ ∇. This proves the inclusion ⊇ in the first
case, the second case as well as the dual proof for µ̃ are analogous.
Next we show that γ̃G is supremum-dense by proving that γ̃G contains all
×
object concepts of B( t∈T Kt ). We have (with g := (gt )t∈T )
Finally, we have
Theorem 32 The congruence lattice of a tensor product of finitely many doubly founded
lattices is isomorphic to the tensor product of the congruence lattices:
Lt ) ∼
O O
C( = C(Lt ).
t∈T t∈T
gt = g1 . m1 - g2 . . . . - gn . mn = mt .
If such a sequence of elements of length n exists, then it also exists for every
number which is larger than n, since in this case there is an attribute k ∈ Lt with
. k and consequently gt . k - gt , whereby the sequence can be extended
gt %
arbitrarily. We can apply Proposition 92 and, since T is finite, we obtain
. m in
g. ×K
t∈T
t ⇔ . mt in Kt for all t ∈ T,
gt .
and consequently
g.
. \ mt for some t ∈ T.
\ m ⇔ gt .
.
4.4 Tensorial decompositions 203
Therefore,
(G, M, .
.
\ )=( ×(G , M , ..\)),
t∈T
t t
The tensor product has been defined as the concept lattice of the context
εs (xs ) := ({g ∈ G | gs ≤ xs } ∪ M ∇ , {m ∈ M | xs ≤ ms } ∪ G∇ ),
Proof In other words, the proposition claims that the extent of s∈S εs (xs ) is
W
exactly the union of the extents of the εs (xs ), s ∈ S, and dually. This immediately
results from the explicit descriptions of these sets which were given above.
Hence, the sublattices εt (Lt ) are mutually distributive: The supremum resp.
infimum of elements from different εt (Lt ) can be obtained by forming the union
of the extents or intents, respectively. This implies a calculation rule which will
be formulated in the following definition:
Definition 76 We call two subsets X and Y of a complete lattice mutually dis-
tributive if the following inequalities hold for every index set S and for every
pair of sequences (xs )s∈S , (ys )s∈S of elements xs ∈ X, ys ∈ Y :
_ ^ _ _
(xs ∧ ys ) ≥ ( xr ∨ ys ),
s∈S R⊆S r∈R s∈S\R
^ _ ^ ^
(xs ∨ ys ) ≤ ( xr ∧ ys ).
s∈S R⊆S r∈R s∈S\R ♦
204 4 Decompositions of concept lattices
For this purpose, it suffices to prove that every attribute concept z which is ≥
the left side is also ≥ the right side of the inequality. Hence, let z be an attribute
concept and assume that
Rz := {r ∈ S | εi (xr ) ≤ z}.
Then we obviously have r∈Rz εi (xr ) ≤ z and can follow the following chain of
W
inferences:
_
z≥ (εi (xs ) ∧ εj (ys ))
s∈S
⇔ ∀s∈S z ≥ εi (xs ) ∧ εj (ys )
⇔ ∀s∈S z ≥ εi (xs ) or z ≥ εj (ys )
⇔ ∀s∈S\Rz z ≥ εj (ys )
_ _
⇔ z≥ εi (xr ) ∨ εj (ys )
r∈Rz s∈S\Rz
^ _ _
=⇒ z≥ ( εi (xr ) ∨ εj (ys ))
R⊆S r∈R s∈S\R
(gt , mt ) 6∈ It
for all t ∈ T . Moreover, if the Kt satisfy the condition from Theorem 47, there
exist elements ht ∈ Gt as well as nt ∈ Mt with
for every t ∈ T . We set h := (ht )t∈T and n := (nt )t∈T and find (h, m) 6∈ I and
(g, n) 6∈ I. If now k ∈ G \ n0 , i.e., kt ∈ Gt \ n0t for all t ∈ T , then for every t ∈ T it
holds that
kt0 ⊆ ht0
and, according to Proposition 91, consequently
k 0 ⊆ h0 ,
of the tensor product for which the restrictions of the projection mapping
O O
πΘ : Lt → Lt Θ, x 7→ [x]Θ
t∈T t∈T
ϕ : L1 ⊗ L2 → L,
(G, M, ∇) := ×K
t∈T
t
of the respective contexts. The closed relation ∇t always corresponds to the sub-
lattice εt (B(Kt )). In the doubly founded case, we can be sure that a subtensorial
product is always induced by a compatible subcontext (H, N, ∇ ∩ H × N ) of
(G, M, ∇). Such subcontexts are described by the following definition:
Definition 78 A subdirect product of contexts
Kt := (Gt , Mt , It ), t ∈ T,
is a compatible subcontext
(H, N, ∇ ∩ H × N )
of the direct product ×t∈T Kt having the property that for each t ∈ T the sub-
context
(Ht , Nt , It ∩ Ht × Nt )
with
Ht := {ht | h ∈ H} and Nt := {nt | n ∈ N }
is dense in Kt . ♦
Proposition 96 The subdirect products of contexts are precisely the compatible subcon-
texts of subtensorial products.
Proof According to Proposition 44, the condition (here restricted to εt (B(Kt )))
that the map ΠH,N is injective, is equivalent to the fact that the subcontext
(H, N, ∇t ∩ H × N ) is dense in (G, M, ∇t ). This in turn is, according to the same
proposition, equivalent to the fact that
holds for each concept (A, B) of (G, M, ∇t ). If we set At := {gt | g ∈ A}, we rec-
ognize by means of the description of the concepts of (G, M, ∇t ) in Proposition
90 that
4.4 Tensorial decompositions 207
{gt | g ∈ A ∩ H} = At ∩ Ht
and therefore
Consequently,
(A ∩ H)∇t ∇t = A ⇔ (At ∩ Ht )It It = At .
This is again the condition from Proposition 44. Hence, (H, N, ∇t ∩ N × N ) is
dense in (G, M, ∇t ), if and only if (Ht , Nt , It ∩ Ht × Nt ) is dense in Kt .
The restriction of the closed relations ∇t to such a subcontext then yields the
subrelations Jt := ∇t ∩ H × N with
B(H, N, Jt ) ∼
= B(Kt ).
K := (G, M, I)
α : G → H, β:M →N
g It m ⇔ αg ∇t βm. ♦
(g, h) ∈ Θt : ⇔ g It = hIt
(m, n) ∈ Ψt : ⇔ mIt = nIt .
with
([g]Θt , [m]Ψt ) ∈ I¯t : ⇔ (g, m) ∈ It
then is the corresponding clarified context.
Then, we can naturally assign a subcontext of the direct product
×K
t∈T
◦
t =( × G/Θ , × M/Ψ , ∇)
t∈T
t
t∈T
t
The role of the maps α and β from Definition 79 is taken over by the maps ι and
ι, which are defined as follows:
ι:G→ × G/Θ ,
t∈T
t g 7→ ([g]Θt )t∈T
ι:M → × M/Ψ ,
t∈T
t m 7→ ([m]Ψt )t∈T .
are surjective on the factor contexts whereby their images are certainly dense.
Furthermore, we have
(ιG, ιM, ∇ ∩ ιG × ιM )
(h, n) 6∈ ∇ and m∇ ⊆ n∇ .
i.e.,
(βmt )∇t ⊆ (βn)∇t for all t ∈ T,
which yields
mIt t ⊆ nIt t for all t ∈ T.
210 4 Decompositions of concept lattices
Hence, we obtain
¯ ¯
([mt ]Ψt )It ⊆ ([nt ]Ψt )It for all t ∈ T
and with n := ιn
m∇ ⊆ n∇ .
Because of (αg, βn) 6∈ ∇, we have (g, n) 6∈ I and therefore (g, n) 6∈ It for all t ∈ T ,
which yields the statement
(h, n) 6∈ ∇,
which we were still lacking.
Proposition 97 contains a structural description of subdirect products. It is
particularly easy to make use of this fact when we are dealing with a doubly
founded context. We shall explain this in the following theorem. The notations
used for this purpose are to be understood as follows: A family (It )t∈T of subre-
lations of (G, M, I) is called doubly founded if each of the contexts (G, M, It ) is
doubly founded. The arrow relations .t and %t also refer to these contexts.
Theorem 34 A doubly founded family (It )t∈T of subrelations is a subdirect decomposi-
tion of (G, M, I) if and only if:
1. I = t∈T It ,
S
2. if gt .t m for all t ∈ T , then there exists an object h ∈ G with gtIt = hIt for all
t ∈ T,
3. if g %t mt for all t ∈ T , then there exists an attribute n ∈ M with mIt t = nIt for all
t ∈ T.
Proof From Proposition 97 we infer that (It )t∈T is a subdirect decomposition if
and only if the subcontext
(ιG, ιM, ∇ ∩ ιG × ιM )
m ≥ x ∧ y ⇔ m ≥ x or m ≥ y
holds for every -irreducible element m ∈ M (L). Hence, two concepts (A1 , B1 )
V
and (A2 , B2 ) certainly form a weakly distributive pair if A1 ∪ A2 is an extent
and B1 ∪ B2 is an intent. We have seen above that this is always the case for
pairs (x, y) of elements of a tensor product with x ∈ εi (Li ), y ∈ εj (Lj ) and
i 6= j. This implied that those sublattices were mutually distributive. In fact, the
following statement can be shown by means of the same proof as as was given
for Proposition 94:
Proposition 98 If all pairs (xt , yt ), t ∈ T are weakly distributive, then
_ ^ _ _
(xt ∧ yt ) = ( xs ∨ yt )
t∈T S⊆T s∈S t∈T \S
t
f
c @
@
d
@ @
b
@
c @ @ @
@ @ @ @
a
@ @ t. .. .. .. . . . . . . . . . . . .@
.. d @t
c @ cc @ cd .
1 @ 4 @ ..........
.... @
@
@ ...
@ t. . . . . . . . . . . . .. .. .. ..@
@
.. . d
f
@ @
@c @c
2 @ 3 @
@ @
@c @ tf
Figure 4.23 The dotted lines in the diagram on the right link the weakly distributive pairs
of incomparable elements.
Proof From the Propositions 95 and 98 it immediately follows that the condi-
tions specified are sufficient for a subtensorial decomposition. It remains to be
212 4 Decompositions of concept lattices
shown that weak distributivity is necessary as well. For this purpose we use The-
orem 34 for the standard context
Corollary 4 Two doubly founded complete sublattices L1 and L2 whose union generates
a sublattice that is also doubly founded are mutually distributive if and only if every pair
(x1 , x2 ) with x1 ∈ L1 and x2 ∈ L2 is weakly distributive.
I a b c d I1 a b c d I2 a b c d
1 ×× 1 × × .
% % 1 .
% .
% %
.
2 ××× 2 × × × .
% 2 .
% .
% .
%
3 ××× 3 .
% .
% . × 3 .
% × × ×
4 × 4 .
% .
% . × 4 %
. .
% .
%
I1 a, b c d
I2 a b, c, d
1 ×
1, 2, 4
2 × ×
3 ×
3, 4 ×
Figure 4.24 Context for the lattice from Figure 4.23, together with the closed relations for
the sublattices. Below, the clarified contexts.
a, b c d
∇
a
b, c, d a b, c, d a b, c, d
1, 2, 4 × × .
% % ←
1
3 × × %
. × % ×
1, 2, 4 × × × × .
% ←
2
3 × × × × .
% ×
1, 2, 4 .
% . × × ←
3,4
3 .
% × . × × × ←
↑ ↑ ↑ ↑
Figure 4.25 The direct product of the two clarified contexts from Figure 4.24. The context
from Proposition 97 is marked by the arrows on the margin.
f
@
({a, b}, {b, c, d})
f @
@
@ ({c},{b, c, d})
({a, b}, {a}) @ ({d}, {b, c, d})
f @f @f
({1}, {1, 2, 4}) @ @ @
@f @f @ f({d}, {a})
@ @ ({3, 4}, {1, 2, 4})
({c}, {a})
f @ f @ f
({2}, {1, 2, 4}) @ ({1}, {3}) ({3, 4}, {3})
@f
({2}, {3}) @
@f
Figure 4.26 The concept lattice for the context from Figure 4.25 is the tensor product of
the sublattices from Figure 4.23.
on the margin. From the arrow relations we can see that it is arrow closed and
thus compatible.
Finally, Figure 4.26 shows the concept lattice of the context from Figure 4.25,
i.e., the tensor product of the two sublattices forming the subtensorial decom-
position. The lattice we started with can be recognized as an interval below the
largest element; its elements are indeed separated by the projection map. >
214 4 Decompositions of concept lattices
4.1
4.2
This section follows [394]. The decomposing and gluing technique described in
this section was developed and successfully employed by Herrmann [190], an
updated version can be found in Day and Herrmann [87]. Vogt [376] has em-
ployed the technique of atlas-decomposition when investigating the structure of
subgroup lattices of finite Abelian groups.
4.3
The substitution sum and the substitution product were used by Luksch and
Wille [266] for the concept-analytic evaluation of pair comparison tests. They
were formally introduced in [267] and thoroughly examined by Stephan [351],
[350]. From the dissertation of Stephan we have taken in particular Theorem 30
and the preparatory Proposition 89. However, we have introduced a little change
in the notation compared with the literature we quote and now write U (a, b)V ,
where in earlier publications appeared U (b, a)V .
4.4
Tensor products of complete lattices have been introduced in many articles. Our
presentation (Sections 4.4, 5.4) does not claim to be complete, but is meant to
supply the basic knowledge. The definition of the tensor product discussed in
this section has been taken from [396] and the generalization in [405]. Precursors
can be found among other things in Waterman [383], Mowat [289] and Shmuely
[342]. A description of the extents and intents of direct products of contexts (as
Gκ -ideals) can be found in [405]. The significance of this product for category
theory was discussed by Erné [116]. Other sources are Bandelt [22], Raney [310],
Kalmbach [208], and, more recently, Chornomaz [76].
Subtensorial products are treated in [160].
Chapter 5
Constructions of concept lattices
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 215
B. Ganter, R Wille, Formal Concept Analysis, https://doi.org/10.1007/978-3-031-63422-2_6
216 5 Constructions of concept lattices
The map
(A, B) 7→ ((A ∩ Gt , B ∩ Mt ) | t ∈ T )
is a natural isomorphism.
The projection map on B(Kt ) combined with this isomorphism is the map
(A, B) 7→ (A ∩ Gt , B ∩ Mt ).
Dually, the extents of the sum context are precisely the unions of extents of the
summands, which yields the isomorphy we claimed. The statement on the com-
patible subcontexts can be verified by means of Proposition 40.
5.1 Subdirect product constructions 217
Proof On Page 151, while proving Theorem 18, it was shown that g Jrs ◦Jst =
g Jrs Is Jst holds for all g ∈ Gr . This proves that 1. is equivalent to 2. The other
equivalence is dual.
g JJ ∩ Gs = g Jrs Jss .
for every t ∈ T , and consequently g J ⊆ hJ . This, together with the dual argu-
ment, shows that the condition (C) of Proposition 57 is satisfied: J is closed.
The Propositions 99 and 101 can be summarized as follows:
Theorem 37 For a subrelation J ⊆ I in the sum context (G, M, I) := Kt , the
P
t∈T
following statements are equivalent:
1. J is a closed relation and corresponds to a subdirect product of the B(Kt ), t ∈ T .
2. J is a closed relation and It = J ∩ (Gt × Mt ) holds for all t ∈ T .
3. The Jst := J ∩ (Gs × Mt ) are bonds from Ks to Kt with Jtt = It and Jrt ⊆ Jrs ◦Jst
for all r, s, t ∈ T .
Note that the last point of the theorem does not contain the requirement that
J is closed. The latter follows (as in Proposition 101) as a consequence if J is
made up of bonds, as specified above. If the contexts Kt are all reduced, then
(G, M, J) is reduced as well. This follows from Proposition 58.
In order to be able to use the subdirect product as a construction method, we
introduce the following notion:
Definition 80 If P is a set, L is a complete lattice and α : P → L is a map, then
we call (L, α) a (complete) P -lattice if L is generated by {αp | p ∈ P }.
When P := {1, 2, . . . , n} we also speak of a (complete) n-lattice. If (P, ≤) is
an ordered set, then we call (L, α) a (complete) (P, ≤)-lattice if α is furthermore
order-preserving.
5.1 Subdirect product constructions 219
αp := (αt p | t ∈ T ), p ∈ P.
We call this lattice the P -product of the lattices (Lt , αt ). As a symbol for P -
P n
products of two lattices we use × or ×, respectively. ♦
d
@
d @d
1 d @ d @ @ d3
@ @ @
d1 d3
@ @ @ @ @
d1,3 d4 d2 @d @d @d @ @
4 4
× × ∼
=
@ @ @ @ @ @
d d3 d1 @ @d @d @d @ @
2,4 @ @ @ @ @
d2 d4 @d @ @d @ @d
4 @ @ @ 2
@d @ @d
@
@d
If the lattices involved are concept lattices, we use the following obvious
terms: (K, α) is called P -context, if (B(K), α) is a P -lattice. In this case, α maps
the elements of P onto concepts of K; for those images we most often write
(Ap , B p ) := αp. Then we call (B(K), α) “the concept lattice of the P -context
(K, α)”, for short.
Evidently, the P -product is a complete subdirect product, since the canonical
projection πt maps the generating system {αp | p ∈ P } of the P -product onto
{αt p | p ∈ P }, i.e., onto a generating system of Lt . Therefore, πt must be sur-
jective. Hence, according to Theorem 37, to the P -product of P -concept lattices
there corresponds in a natural way a closed relation J in the sum context. This is
220 5 Constructions of concept lattices
described more precisely in the following theorem, in which we shall again use
the abbreviation Jst := J ∩ (Gs × Mt ):
Theorem 38 The closed relation J of the sum context t∈T Kt which belongs to a P -
P
product of P -concept lattices (B(Kt ), αt ) is characterized by the following properties:
1. for all t ∈ T , Jtt = It ,
2. for all s, t ∈ T, s 6= t, Jst is the smallest bond from Ks to Kt which contains the sets
Aps × Btp , p ∈ P,
Aps and Btp being defined to be αs p =: (Aps , Bsp ) and αt p =: (Apt , Btp ).
Proof From the definition of a bond (Definition 61 on p. 148) it immediately fol-
lows that the intersection of bonds is a bond. Therefore, there is always a smallest
bond from Ks to Kt , s 6= t which entirely contains the sets
Aps × Btp , p ∈ P.
is precisely the relation which is characterized by the two conditions of the the-
orem. We shall show first that J satisfies condition 3) of Theorem 37.
For this purpose we consider, for fixed r, s, t ∈ T , p ∈ P , an attribute m ∈ Btp .
Because of Aps × Btp ⊆ Jst we certainly have mJst ⊇ Aps . Likewise, because of
Apr × Bsp ⊆ Jrs , we can infer for each object g ∈ Apr that g Jrs ⊇ Bsp and, since
(Aps , Bsp ) is a concept of Ks , even g Jrs Is ⊆ Aps . Hence, we have g Jrs Is ⊆ Aps ⊆ mJst
and obtain, together with Definition 62 on p. 150,
Since this is correct for all g ∈ Apr and all m ∈ Btp , we have
Theorem 18 on p. 151 furthermore states that Jrs ◦ Jst is a bond. Since we have
assumed that Jrt is the smallest bond containing all those sets, it follows that
J contains all sets Ap × B p , and every closed relation containing those sets be-
longs to a subdirect product and therefore has to satisfy the third condition of
Theorem 37. J is the smallest closed relation for which this is true, and is there-
fore, according to Proposition 55, the closed relation belonging to the sublattice
generated by (Ap , B p ).
˙
[ ˙
[
αp := ( Atp , Btp ).
t∈T t∈T
In this case, the method used in the preceding example would lead to compli-
cated intermediate steps. Therefore, we determine the corresponding 4-standard
contexts and obtain:
222 5 Constructions of concept lattices
bg
4
b
m1 m1 m2 m3 m4
m2 b @ bm3 g1 × × × α1 (1) = α1 (2) = ({g1 , g2 }, {m1 , m2 })
g2 g3 g2 × ×
@b α1 (3) = α1 (4) = ({g1 , g3 }, {m1 , m3 })
g1 g3 × ×
m4 b g4
b
n1 n2 n3 n4 n5 α2 (1) = ({h1 }, {n1 , n2 , n3 , n5 })
n2 b @ b n1
h4 h5 h1 × × × × α2 (2) = ({h1 , h2 , h3 , h4 }, {n2 })
@b h2 × × × ×
h3
n3 b h3 × × α2 (3) = ({h2 }, {n1 , n2 , n3 , n4 })
n5 b @ b n4 h4 ×
h1 h2 α2 (4) = ({h1 , h2 , h3 , h5 }, {n1 })
h5 ×
@b
Now we form the 4-fusion of these two contexts, as described in Theorem 38,
i.e., we form the disjoint union of the two contexts and add the sets {g1 , g2 } ×
{n1 , n2 , n3 , n5 }, {g1 , g2 } × {n2 }, {g1 , g3 } × {n1 , n2 , n3 , n4 }, {g1 , g3 } × {n1 } for J1,2
as well as the sets {h1 } × {m1 , m2 }, {h1 , h2 , h3 , h4 } × {m1 , m2 }, {h2 } × {m1 , m3 },
{h1 , h2 , h3 , h5 } × {m1 , m3 } for J2,1 to the incidence. In the present case this has
already resulted in bonds. In general, the incidence must be extended until bonds
are obtained. As a result we obtain the following 4-context:
m 1 m 2 m 3 m 4 n1 n2 n3 n4 n5
g1 × × × × × × × ×
g2 × × × × × × α(1) = ({g1 , g2 , h1 }, {m1 , m2 , n1 , n2 , n3 , n5 })
g3 × × × × × × α(2) = ({g1 , g2 , h1 , h2 , h3 , h4 }, {m1 , m2 , n2 })
g4
h1 × × × × × × × α(3) = ({g1 , g3 , h2 }, {m1 , m3 , n1 , n2 , n3 , n4 })
h2 × × × × × × ×
α(4) = ({g1 , g3 , h1 , h2 , h3 , h5 }, {m1 , m3 , n1 })
h3 × × × × ×
h4 × × ×
h5 × × ×
αt g := (g tt , g t ) ∈ B(Kt ) for g ∈ G,
t tt
αt m := (m , m ) ∈ B(Kt ) for m ∈ M.
d
d
@
d d d @d
@ @ @
d 2 d
@ d4 2 d
@d @ d4
@ @ @ @
1,2 d @ d3,4 4 @d ∼ @d d @d
× =
@ @ @
@d d d @d @d
@ @ @
d 1 d
@ d3 1 d
@d @ d3
@ @ @
@d @d @d
@
@d
Figure 5.2 The 4-lattice on the right is the concept lattice of the 4-fusion calculated in
Example 5.1 and is consequently isomorphic to the 4-product of the factors on the left.
5.2 Gluings
I ⊆ G × M0 ∪ G0 × M.
M0
Proof “⇒”: If (g, m) ∈ I, then (M00 , M0 ) ≤ (g 00 , g 0 ) (i.e., m ∈ G0
M0 ) or (g 00 , g 0 ) ≤ (G0 , G00 ) (i.e., g ∈ G0 ). “⇐”: If I ⊆ G ×
M0 ∪ G0 × M and (X, Y ) is a concept with X 6⊆ G0 , then ∅
X 0 ⊆ M0 and therefore (M00 , M0 ) ≤ (X, Y ).
224 5 Constructions of concept lattices
If we are confronted with the situation described in the proposition, the con-
cept lattice is made up in a simple way of two lattices, namely of the ideal
((G0 , G00 )] and the filter [(M00 , M0 )), which overlap in the (possibly empty) in-
terval [(M00 , M0 ), (G0 , G00 )]. We speak of the Hall-Dilworth gluing (cf. Defini-
tion 70 on page 175), in the special case (G0 , G00 ) = (M00 , M0 ) of the vertical sum,
which has already been mentioned on page 42. There we also introduced the
horizontal sum, where two lattices are “glued together sideways” by identify-
ing the two largest and the two smallest elements. In Section 4.3 we showed that
the substitution product generalizes both constructions.
Whereas the Hall-Dilworth gluing is the simplest case of an atlas construction
in the sense of Section 4.2, another way of generalizing the horizontal sum sug-
gests itself, namely the horizontal gluing, where we allow the lattices involved to
overlap in more than the two border elements. The general situation, namely the
situation that a concept lattice is the union of sublattices, is treated in the next
(rather trivial) proposition:
Proposition 103 For relations Jt ⊆ G × M , t ∈ T , the following statements are equiv-
alent:
1. B(G, M, t∈T Jt ) = t∈T B(G, M,
S S
SJt ).
2. The Jt are closed relations of (G, M, t∈T Jt ) and
[
A×B ⊆ Jt ⇒ ∃s∈T A × B ⊆ Js .
t∈T
The proof is simple, but the result is not very rewarding. In order to obtain a con-
dition which is easier to manage, we limit ourselves to two lattices and assume
that the overlapping is the union of an ideal and a filter, i.e., that it has the form
described in Proposition 102.
Definition 83 A complete lattice L is an ideal-filter gluing of two sublattices U1
and U2 if:
1. L = U1 ∪ U2
2. x ≤ y in L implies {x, y} ⊆ U1 or {x, y} ⊆ U2 .
3. U1 ∩ U2 = (a] ∪ [b) for suitable elements a, b ∈ L. ♦
An ideal-filter gluing can be recognized by the context:
Theorem 39 The following conditions are equivalent:
1. B(G, M, I) is an ideal-filter gluing of complete sublattices U1 and U2 . J1 := C(U1 )
and J2 := C(U2 ) are the corresponding closed relations.
2. J1 and J2 are closed relations of (G, M, I) and
a. g I = g J1 or g I = g J2 holds for every object g and, dually, mI = mJ1 or mI =
mJ2 holds for every attribute m,
b. there is an extent G0 and an intent M0 of (G, M, I) with J1 ∩ J2 = (G0 × M ∪
G × M0 ) ∩ I.
5.2 Gluings 225
M0
G0
J1 ∅
∅ J2
Figure 5.3 Context of an ideal-filter-gluing. In the hatched area J1 and J2 coincide with
I ; this is at the same time the closed relation belonging to the overlapping of the lattices.
Proof 1) ⇒ 2a): Every object concept γg belongs to one of the sublattices, but
γg ∈ Ui is equivalent to g I = g Ji .
1) ⇒ 2b): By assumption U1 ∩ U2 = (a] ∪ [b) for suitable concepts a =: (G0 , G00 )
and b =: (M00 , M0 ). It is evident that J1 ∩J2 ⊇ (G0 ×M ∪G×M0 )∩I, what remains
to be shown is the other inclusion. Assume that (g, m) ∈ J1 ∩ J2 and m 6∈ M0 .
Then there are concepts (X1 , Y1 ) ∈ U1 and (X2 , Y2 ) ∈ U2 with (g, m) ∈ Xi × Yi .
The infimum of these concepts is comparable with both of them, i.e., one of the
three concepts (X1 , Y1 ), (X2 , Y2 ) and (X1 , Y1 ) ∧ (X2 , Y2 ) is not in U1 ∩ U2 . The
intent of this concept contains m, i.e., it cannot be a subset of M0 . Hence, the
concept is contained in the ideal ((G0 , G00 )], and we obtain X1 ∩ X2 ⊆ G0 , which
implies g ∈ G0 .
2) ⇒ 1): We have to show that all concepts and all comparabilities between
concepts of (G, M, I) originate from one of the sublattices Ui . Let us assume
that there is a concept (X, Y ) ∈ B(G, M, I) that belongs neither to U1 nor to
U2 , so that neither X × Y ⊆ J1 nor X × Y ⊆ J2 . Then there must be pairs
(g, m), (h, n) ∈ X × Y with (g, m) ∈ J1 \ J2 and (h, n) ∈ J2 \ J1 , and from the pre-
suppositions it follows that g, h 6∈ G0 and m, n 6∈ M0 . Hence, (g, n) cannot belong
to J1 ∩ J2 , but certainly (g, n) ∈ I, since (g, n) ∈ X × Y . From (g, m) ∈ J1 \ J2 we
infer g I = g J1 , i.e., (g, n) ∈ J1 , and from (h, n) ∈ J2 \ J1 we dually infer nI = nJ2 ,
i.e., (g, n) ∈ J2 , which is a contradiction.
If (X, Y ) ∈ B(G, M, J1 ) and (U, V ) ∈ B(G, M, J2 ) and (U, V ) ≤ (X, Y ), then
U ⊆ X and V ⊆ Y , i.e., U × Y ⊆ J1 ∩ J2 . This implies U ⊆ G0 (i.e.,
(U, V ) ∈ B(G, M, J1 )) or Y ⊆ M0 (i.e., (X, Y ) ∈ B(G, M, J2 )), in any case
{(U, V ), (X, Y )} ⊆ Ui for i = 1 or i = 2.
The characterization in Theorem 39 leads the way to the corresponding con-
text construction. The following definition is illustrated by Figure 5.4:
Definition 84 The union of two formal contexts K1 := (G1 , M1 , I1 ) and K2 :=
(G2 , M2 , I2 ) is the context
226 5 Constructions of concept lattices
K1 ∪ K2 := (G1 ∪ G2 , M1 ∪ M2 , I1 ∪ I2 ).
M0
∅
K1
G0
K2
∅
The context gluing is not the exact counterpart of the lattice gluing. The pre-
conditions are weaker. The condition that K1 and K2 coincide on G0 × M0 , does
not at all enforce that the extents contained in G0 and the intents contained in
M0 are also the same in both contexts. This is however necessarily true in the
case of an ideal-filter gluing. Therefore, it is rather surprising that the following
theorem holds true. There is a snag in the theorem, which we have to point out.
It says that the concept lattice of the gluing of two contexts K1 and K2 is the
ideal-filter gluing of two sublattices, but it does not say that those sublattices are
isomorphic to B(K1 ) and B(K2 ). In fact, this is generally not the case.
Theorem 40 B(K) is the ideal-filter gluing of two complete sublattices U1 and U2 with
G0 = G1 ∩ G2 and M0 = M1 ∩ M2 .
G1 := {g ∈ G | g I = g J1 },
5.2 Gluings 227
M1 := {m ∈ M | mI = mJ1 },
G2 := G0 ∪ (G \ G1 ),
M2 := M0 ∪ (M \ M1 )
and I1 := I ∩ G1 × M1 ,
I2 := I ∩ G2 × M2
J1 := I1 ∪ I ∩ (G0 × M ∪ G × M0 )
J2 := I2 ∪ I ∩ (G0 × M ∪ G × M0 ).
Then, we have to prove that J1 and J2 are closed relations. We show this for
J := J1 , with the help of Proposition 56: Let X ⊆ G be arbitrary. If X 6⊆ G0 , then
X J ⊆ M1 and because of
J ∩ G × M1 = I ∩ G × M1
T is extent of K1 ⇔ T is extent of K2
and the corresponding is true for M0 . Those conditions have the effect that the
ideals generated by the concept with the extent G0 are isomorphic in B(K1 ) and
B(K1 ) and that the same is true for the filters of the concepts with the intents
⊆ M0 . The fact that K1 and K2 coincide in I0 implies that those isomorphisms
can be generated by a single map.
In particular, we have: An ideal-filter gluing of two concept lattices is isomor-
phic to the concept lattice of the gluing of the contexts involved.
In practice, the task that usually crops up is the slightly generalized one of
having to glue two lattices together which do not have elements in common, but
in which an isomorphism of the ideal-filter pair of one lattice onto a correspond-
ing pair of the other lattice is given. In order to implement this construction for
concept lattices, one first modifies the respective contexts K1 and K2 in such a
228 5 Constructions of concept lattices
way that both the object concepts in the two ideals and the attribute concepts in
the two filters coincide. This can be achieved through mutual enrichment, and
in the case of doubly founded contexts even through reduction. The objects and
attributes of these concepts are given the same name, if they are mapped onto
each other by the isomorphism. For the remainder one makes the two contexts
disjoint. The concept lattice of the gluing of the contexts modified in this way is
then the ideal-filter gluing of suitable isomorphic copies of their concept lattices,
as desired.
e.......... e e
H H
@ HH @ HH
b e ec a
e.......... a e @c e H ed b e e c ea @ ec H ed
HH HH HH HH
@ H
@ @ @ −→ @ H@ @
@
e e e .......... e 3 e H@ e4 e e @ e@ He 3H
@e
H
2 1H
2 H 1
4
@ H@ @ @
0
@ 0 0
HH HH
@
H@ e.......... e
HH @ e
a b c d e
a b c a d e 0 × × × ×
0 × × 0 × × 1 × ×
1 × × ∪ 3 × × = 2 × ×
2 × × 4 × × 3 × ×
4 × ×
K1 K2
K1 ∪ K2
Hence, under the conditions of the proposition, the map ΠG,M with
(A, B) 7→ (A ∩ G, B ∩ M )
Thus it is only in the first case that (C, D) has two different pre-images with
respect to ΠG,M . Additionally, we note down:
Proposition 106 If (G ∪ H, M ∪ N, J) is a context with the properties specified in the
preceding proposition, then:
(A, B) is a concept of (G ∪ H, M ∪ N, J) if and only if (A ∩ G, B ∩ M ) is a concept of
(G, M, I) and we are dealing with one of the following three cases
1. A ⊆ G, B ⊆ M , A = B J , B = AJ
2. A ⊆ G, B = AJ 6⊆ M
3. B ⊆ M, A = B J 6⊆ G.
Proof According to Proposition 104, for every concept (A, B) of (G ∪ H, M ∪
N, J), the restriction (A ∩ G, B ∩ M ) is a concept of (G, M, I), and, since we have
presupposed that J ∩ H × N = ∅, it follows that A × B ∩ H × N = ∅, i.e., A ⊆ G
or B ⊆ N , and thus one of the cases 1)–3) must hold.
If, conversely, A is an extent of (G, M, I) and AJ 6⊆ M , then (A, AJ ) by the
preceding proposition is a concept of (G∪H, M ∪N, J). If AJ = AI , then B = AJ
is an intent of (G, M, I), and, under the condition that B J 6= B I , we can argue
dually. What remains is the trivial case AI = AJ and B I = B J .
3. g % n, n ∈ N together imply g ∈ H.
We have made no restrictions concerning the choice of the sets H and N . How-
ever, it turns out that we can make a very special choice without loss of generality.
Definition 85 Assume that C ⊆ B(G, M, I) is a convex set of concepts and
w.l.o.g. that C ∩ (G ∪ M ) = ∅. Then
K[C] := (G ∪ C, M ∪ C, IC ), ♦
IC ∩ G × M := I, IC ∩ C × C := ∅
g IC (C, D) : ⇔ g ∈ C,
(C, D) IC m : ⇔ m ∈ D.
Evidently, the context defined in this way satisfies the conditions from Propo-
sition 104, C assuming the role of H as well as of N .
M N M C
G I G I ∈
H ∅ C
∈ ∅
(G ∪ H, M ∪ N, J) K [C ]
Figure 5.7 In the case of the doubling construction H and N can be replaced by the same
convex set C.
γI (h)IC = hJ ,
µI (n)IC = nJ .
B ⊆ hJ ⇔ B ⊆ γI (h)IC
and dually. Since furthermore for every concept (C, D) ∈ C by the definition of C
there exists some h ∈ H and some n ∈ N with µI (n) ≤ (C, D) ≤ γI (h), we have
AJ ∩ H 6= ∅ ⇔ AIC ∩ C 6= ∅ and B J ∩ N 6= ∅ ⇔ B IC ∩ C 6= ∅.
(A, AIC ) if A ⊆ G
ϕ(A, B) := ,
(B IC , B) if B ⊆ M
ϕ : B(G ∪ H, M ∪ N, J) → B(K[C]).
First of all, we note that ϕ(A, B) by Proposition 106 is defined for every concept
(A, B) ∈ B(G ∪ H, M ∪ N, J). By means of the equivalence proved above and
again by means of Proposition 106 we conclude that ϕ(A, B) is in fact always
a concept of B(K[C]) and even that every such concept occurs. Hence, ϕ is a
bijection. The fact that ϕ is also an order isomorphism is elementary because of
the simple shape of the concepts involved.
Thanks to Proposition 107 we may concentrate on the context construction
K 7→ K[C], because it covers the general case. The content of Proposition 105,
specialized to the context K[C], reads as follows:
Proposition 109 For every concept (C, D) of K there is at least one and at most two
concepts (A, B) of K[C] with (A ∩ G, B ∩ M ) = (C, D), namely
5.3 Local doubling 233
(C, C IC ) and (DIC , D) are both concepts of K[C] and distinct from it if and only if
(C, D) ∈ C.
Proof What remains to be proved is only the last sentence. By Proposition 105
there are two concepts (A, B) with (A ∩ G, B ∩ M ) = (C, D) if and only if there
are elements h, n ∈ C with C ⊆ hIC , D ⊆ nIC , i.e., h ≤ (C, D) ≤ n. Since C is
convex, this is equivalent to (C, D) ∈ C.
Definition 86 For a convex subset C of a complete lattice L we define the com-
plete lattice L[C] := (L[C], ≤) to be
and
x, y ∈ L \ C and x ≤ y in L
or x ∈ L \ C, y = (y0 , i), y0 ∈ C, x ≤ y0 in L
x ≤ y :⇔
or y ∈ L \ C, x = (x0 , i), x0 ∈ C, x0 ≤ y in L
or x = (x0 , i), y = (y0 , j) ∈ C × {0, 1}, i ≤ j and x0 ≤ y0 . ♦
The assertion that a complete lattice is defined in this way requires a proof. It
follows from the next theorem.
B(K)[C] ∼
= B(K[C]).
if (A ∩ G, B ∩ M ) 6∈ C,
(A ∩ G, B ∩ M ),
ϕ(A, B) := ((A ∩ G, B ∩ M ), 0), if (A, B ∩ M ) ∈ C,
((A ∩ G, B ∩ M ), 1), if (A ∩ G, B) ∈ C,
defines an isomorphism
ϕ : B(K[C]) → B(K)[C].
⇔ A1 ∩ G ⊂ A2 ∩ G or (A1 ∩ G, B1 ∩ M ) ∈ C
⇔ ϕ(A1 , B1 ) < ϕ(A2 , B2 ).
In practice, we would if possible reduce the context K[C]. A look at the arrow
relations shows us how: If c ∈ C is an object of K[C], then by Proposition 107
c . d can only hold for one attribute d ∈ C. We discover quickly that precisely
the minimal resp. maximal elements of C are irreducible. If we define
Cmin := {c ∈ C | c is minimal in C}
c . d ⇔ c ∈ Cmin and c ≤ d,
c % d ⇔ d ∈ Cmax and c ≤ d.
If, therefore, we assume that C has enough minimal and maximal elements, i.e.,
that [
C = {[c, d] | c ∈ Cmin , d ∈ Cmax , c ≤ d},
then K[C] is doubly founded (provided that K is doubly founded), and the con-
text K[C] has (up to isomorphism) the same concept lattice as
Particularly simple is K[C]r in the case of the interval doubling, that is in the
case that C is an interval C = [(B 0 , B), (C, C 0 )] of B(K), because in this case Cmin
and Cmax are both one-element (sets) and we have
with
J ∩ G × M = I, (B 0 , B)J := B and (C, C 0 )J := C.
We note this down as a proposition:
Proposition 110 A doubly founded context is of the form K[C] for an interval C ⊆
B(K), if and only if there is an object h and an attribute n with
h%
. n, g . n ⇒ g = h, h % m ⇒ m = n.
Because in this case with H := {h} and N := {n} the conditions of Proposition
107 are evidently satisfied.
Example We consider the possible bracketings of a product x0 x1 · · · xn of n+1
variables x0 , . . . , xn . Since the names of the variables are of no consequence, we
replace them by dots. Thus, (..)((..).) stands for (x0 x1 )((x2 x3 )x4 ), etc. We can
5.3 Local doubling 235
order these bracketings by agreeing that a term becomes larger if subterms are
replaced according to the rule
A(BC) −→ (AB)C.
Tamari [367] observed that this induces an order which turns the set of all brack-
etings of n + 1 symbols into a lattice; this lattice is therefore called the Tamari
lattice Tn . Bennett and Birkhoff [43] have determined the irreducibles of these
lattices. This makes it possible to state a (reduced) context for the Tamari lattice
Tn . With S := {1, 2, . . . , n} and P2 (S) := {{i, j} | i, j ∈ S, i 6= j} this is the
context
(P2 (S), P2 (S), I),
the incidence I for i < j and p < q being defined by
Geyer [171] has stated a recursion rule for these contexts, which can be recog-
nized by means of the example n = 5 in Figure 5.8.
1 2 1 3 2 1 4 3 2 1
2 3 3 4 4 4 5 5 5 5
1 2 %
. × × × × × × × × ×
2 3 × .
% . × × × × × × ×
1 3 % × .
% × × × × × × ×
3 4 × × × .
% . . × × × ×
2 4 × % × .
% . × × × ×
1 4 % × % × × .
% × × × ×
4 5 × × × × × × .
% . . .
3 5 × × × % × .
% . .
2 5 × % × % × × .
% .
1 5 % × % × × % × × × .
%
The context shows a salient structure of the arrow relations: The (square)
cross table can be arranged in such a way that all double arrows are on the main
diagonal, all upward arrows are above and all downward arrows are below the
main diagonal. In this particular case the “lowest” object h and the correspond-
ing attribute with regard to %
. n evidently satisfy the conditions 110. This means
that the context is generated by interval doubling from the subcontext obtained
by omitting h and n.
236 5 Constructions of concept lattices
However, this subcontext has again the same structure of the arrow relations.
Hence, the procedure can be repeated until there remains nothing. This means
that the Tamari lattice can be generated by iterated interval doubling from the
one-element lattice. Figure 5.9 shows the Tamari lattice T4 including its “gene-
(((..).).).
e
4
@
@
((.(..)).). e
@
@
5 @ e((..)(..)). @
@ @
(.((..).)). e
@ @
@ @
4
6
@ @
e @ @ e((..).)(..)
.(((..).).) e @ @
@ e (..)((..).)
@ @ 4
(.(.(..))). @ @@
2
@ e
(.(..))(..)
@ @
4
6 @ @
.((.(..)).) e @ @
@ @
3 5
1
@ @
@
@ @ e(..)(.(..))
@
.(.((..).)) e @ e.((..)(..))
@
@
@ 1 4
@ 2
@
@
@
@
@
@e
.(.(.(..)))
Figure 5.9 The Tamari lattice T4 . The numbers at the edges indicate the recursive struc-
ture of the lattice, which was generated by interval doubling.
sis”: At the edges, we have noted at which stage of the iterated interval doubling
they have been generated. In descending order, congruences arise, which grad-
ually factorize the lattice until a one-element lattice is reached. >
5.4 Tensorial constructions 237
ε1 : L1 ,→ L1 ⊗ L2 , ε2 : L2 ,→ L1 ⊗ L2
α1 : L1 ,→ L and α2 : L2 ,→ L
such that the complete sublattices α1 (L1 ) and α2 (L2 ) are mutually distributive,
then there is a complete homomorphism
ϕ : L1 ⊗ L2 → L
with
α 1 = ϕ ◦ ε1 , and α 2 = ϕ ◦ ε2 .
⊗4) The union ε1 (L1 ) ∪ ε2 (L2 ) of the two sublattices generates L1 ⊗ L2 .
By these properties the tensor product is characterized up to isomorphism.
Proof The properties ⊗1), ⊗2) and ⊗4) have already been proved. We can easily
see that the properties ⊗1)–⊗4) are characteristic, since for every lattice with
these properties, from ⊗3) we immediately obtain an isomorphism to the tensor
product.
What remains to be shown is ⊗3). Hence, let L be a lattice with the properties
specified in ⊗3). First, we work out the following sub-claim:
For every subset X ⊆ L1 × L2 we have
_ ^
(α1 (x1 ) ∧ α2 (x2 )) = (α1 (y1 ) ∨ α2 (y2 )).
(x1 ,x2 )∈X (y1 ,y2 )∈X ∇
For this purpose, we make use of the condition that the two image sets are mu-
tually distributive and obtain:
_ ^ _ _
(α1 (x1 ) ∧ α2 (x2 )) = ( α1 (x1 ) ∨ α2 (x2 ))
(x1 ,x2 )∈X R⊆X x1 ∈R x2 ∈X\R
238 5 Constructions of concept lattices
^ _ _
= (α1 ( x1 ) ∨ α2 ( x2 )).
R⊆X x1 ∈R x2 ∈X\R
as desired.
Finally, we have to examine the connection between the maps αi and εi . For
this purpose we recall the definition of the εi (in particular of ε1 ), from which it
follows that, for an arbitrary x ∈ L1 , the extent of ε1 (x) is given by {(x1 , x2 ) ∈
L1 × L2 | x1 ≤ x or x2 = 0}. Thereby we obtain
_ _
ϕ(ε1 (x)) = (α1 (x1 ) ∧ α2 (x2 )) ∨ (α1 (x1 ) ∧ α2 (x2 ))
x1 ≤x x1 ≤1
x2 ≤1 x2 ≤0
and µ(x1 , x2 ) =
and therefore
∧ : L1 × L2 → L1 ⊗ L2 and ∨ : L1 × L2 → L1 ⊗ L2
are defined by
x1 ∧ x2 := γ(x1 , x2 ) = ε1 (x1 ) ∧ ε2 (x2 ),
x1 ∨ x2 := µ(x1 , x2 ) = ε1 (x1 ) ∨ ε2 (x2 ). ♦
Proposition 111 The tensorial operations satisfy the following arithmetic rules:
x1 ∧ 1 = ε1 (x1 ) = x1 ∨ 0, 1 ∧ x2 = ε2 (x2 ) = 0 ∨ x2 ,
x1 ∧ x2 = x1 ∨ 0 ∧ 0 ∨ x2 , x 1 ∨ x2 = x1 ∧ 1 ∨ 1 ∧ x2 ,
^ ^ ^ _ _ _
x1s ∧ xs2 = ( xs1 ) ∧ ( x2s ), x1s ∨ xs2 = ( xs1 ) ∨ ( xs2 ),
s∈S s∈S s∈S s∈S s∈S s∈S
_ _ ^ ^
x1 ∧ x2s = x1 ∧ ( xs2 ), x1 ∨ xs2 = x1 ∨ ( xs2 ),
s∈S s∈S s∈S s∈S
_ _ ^ ^
x1s ∧ x2 = ( x1s ) ∧ x2 , xs1 ∨ x2 = ( xs1 ) ∨ x2 ,
s∈S s∈S s∈S s∈S
240 5 Constructions of concept lattices
1∨ 0=1∧ 1=0∨ 1 (= 1 ∨ x = 1 ∨ y = 1 ∨ z
e = w ∨ 1 = 1 ∨ 1)
@
@
@
w∨x w y
e ∨ @ @e ∨ w z
e
@
@ @
@
@ @
@ @ @
0∨x
e @ e0 ∨ y @
@ e w ∨ 0@
w ∧ 1@ @ e0 ∨ z
1∧x @ @ 1∧y 1∧z
@ @
@ @
@@e @e
ew ∧ y @
w∧x @ w∧z
@
@
@e
@
= w ∧ 0 = 0 ∧ 0)
0∧ 1=0∨ 0=1∧ 0 (= 0 ∧ x = 0 ∧ y = 0 ∧ z
1
1 c c
Figure 5.10 w c ⊗ x c c@
y cz
@ with the tensorial operations.
0 c
@@c
0
^ _ ^ ^
(xs1 ∨ xs2 ) = (( xr1 ∧ 1) ∧ ( 1 ∧ xs2 )),
s∈S R⊆S r∈R s∈S\R
_ ^ _ _
(xs1 ∧ xs2 ) = (( xr1 ∨ 0) ∨ ( 0 ∨ xs2 )).
s∈S R⊆S r∈R s∈S\R
Proof All these rules result immediately from the definitions, apart from the last
two, for which we have to consult Proposition 94: Because of x1 ∧ x2 = ε1 (x1 ) ∧
ε2 (x2 ), x1 ∨ x2 = ε1 (x1 ) ∨ ε2 (x2 ) and the rules mentioned in the first line, the
equations are precisely the translation of the circumstance that the sublattices
ε1 (L1 ) and ε2 (L2 ) are mutually distributive.
Thiele [369] has impressively demonstrated how to obtain readable diagrams
of tensor products of small lattices. First, the idea of the P -product developed in
Section 5.1 is transferred to the tensor product and it is agreed that:
Definition 88 For a P -lattice (L1 , α1 ) and a Q-lattice (L2 , α2 ) with P ∩ Q = ∅,
5.4 Tensorial constructions 241
α : P ∪˙ Q → L1 ⊗ L2
is defined as follows:
if r ∈ P ,
ε1 α1 (r)
α(r) :=
ε2 α2 (r) if r ∈ Q.
♦
By means of Theorem 42 we quickly convince ourselves of the fact that this
indeed defines a P ∪˙ Q-lattice.
If L1 and L2 are concept lattices, we can introduce the corresponding context
operation:
Definition 89 For a P -context (K1 , α1 ) and a Q-context (K2 , α2 ) with P ∩ Q = ∅
we define
(K1 , α1 ) × (K2 , α2 ) := (K1 × K2 , α)
to be the P ∪˙ Q-context for which the map
α : P ∪˙ Q → B(K1 × K2 )
B((K1 , α1 ) × (K2 , α2 )) ∼
= (B(K1 ), α1 ) ⊗ (B(K2 ), α2 ).
Thiele has shown that the product defined in this way is distributive over the
P -fusion, i.e., that it is possible to transfer Proposition 18 to this case. This is the
content of the following theorem.
Theorem 43 If (K1 , α1 ) and (K2 , α2 ) are both P -contexts and if (K3 , α3 ) is a Q-context
with P ∩ Q = ∅, then
P
((K1 , α1 ) + (K2 , α2 )) × (K3 , α3 )
˙
P ∪Q
= ((K1 , α1 ) × (K3 , α3 )) + ((K2 , α2 ) × (K3 , α3 )).
Proof Both sides of the equation claimed describe closed relations of
K := (K1 + K2 ) × K3 = K1 × K3 + K2 × K3
(cf. Proposition 18). If we are able to show that the map α is also the same in both
cases, nothing remains to be proved, since in this case the sublattices generated
by
242 5 Constructions of concept lattices
{αx | x ∈ P ∪˙ Q}
and thus the corresponding closed relations must also be the same. This can be
checked easily; the main problem is that of a transparent notation. We again use
the abbreviations
Then we have
α12 p = α1 p + α2 p.
For the embedding maps we use the symbols ε1 , ε2 , ε12 and ε3 in the obvious
way, in the case of ε3 , however, we have to differentiate: We write εi3 , if we are
working in the product Ki × K3 . For reasons of readability we presuppose that
none of the contexts contains any full columns or full rows. Thus, the trivials
terms M ∇ and G∇ disappear when we evaluate the maps εi by means of the
formula stated in Proposition 90.
For the left-hand side we obtain, if p ∈ P ,
αp = ε12 α12 p
= ε12 (A1p ∪ Ap2 , B1p ∪ B2p )
= ((Ap1 ∪ A2p ) × G3 , (B1p ∪ B2p ) × M3 )
and for q ∈ Q
αq = ε12
3 α3 q
= ε312 (A3q , B3q )
= ((G1 ∪ G2 ) × A3q , (M1 ∪ M2 ) × B3q ).
αp = ε1 α1 p + ε2 α2 p
= (A1p × G3 , B1p × M3 ) + (Ap2 × G3 , B2p × M3 )
= ((A1p ∪ Ap2 ) × G3 , (B1p ∪ B2p ) × M3 )
and for q ∈ Q
αq = ε13 α3 q + ε23 α3 q
= (G1 × Aq3 , M1 × B3q ) + (G2 × A3q , M2 × B3q )
= ((G1 ∪ G2 ) × Aq3 , (M1 ∪ M2 ) × B3q ).
5.4 Tensorial constructions 243
M1 M2 c
c @ c3
× × ×
c1 c3@ c2
@ c
1
G1 × G2 × c2
\
× @@c × \c
M1 M2 M1 M2 M1 M1 M2 M2
× ×××× ××× × × ××××××
G1 × ×× × ×× G1 × × ×× ××
××× ××× × ××× ××
××××× ××××× × ×××××××××
G2 ×××× ×××× G1 × ××××× ×××
× ×× × = ×××××× ×××
× ××××××××× ×××××××× ××
G1 × ××××××××× G2 ××××××× ×
××× ×××××× × × × ×
××××× ×××××× ×××××××× ×××
G2 ×××× ×××××× G2 ××××××× ×××
× ××××××× × ××× ××××
Corollary 7 For two any P -lattices (L1 , α1 ), (L2 , α2 ) and a Q-lattice (L3 , α3 ) (with
P ∩ Q = ∅) we have
244 5 Constructions of concept lattices
P
((L1 , α1 ) × (L2 , α2 )) ⊗ (L3 , α3 )
P ∪Q
∼ ((L1 , α1 ) ⊗ (L3 , α3 )) × (L2 , α2 ) ⊗ (L3 , α3 )).
=
This corollary also goes back to Thiele. He has used it skillfully in order to
draw diagrams of tensor products. As a first application we calculate the tensor
product of two four-element chains.
Example In order to calculate the tensor product of two four-element chains,
we make use of the fact that such a chain can be written as a 2-product of a two-
element and a three-element chain:
c c
c
c2 2 c @ 2 c2
= c 1,2 × c .
@
∼
c1 = @c
1
1 c
c @c
@
Using Corollary 7 this can be multiplied out. The convention “tensor product
first, then the P -product” saves brackets. The above expression yields
c c c c
c 4 c2 c2 c4
c 1,2 ⊗ 4 ×
4 4
∼ c 1,2 ⊗ c 3,4 × c
= c3 c 1 ⊗ 3,4 × c1 ⊗ c3 .
c c c c
As can be easily seen by means of the contexts, the tensor product of two three-
element chains has six elements. With respect to the tensor product, two-element
chains behave like neutral elements. Using these observations, we can convert the
expression into a mere 4-product:
c
c c4 c2
c 2,4
∼ 1,2 c @ c3,4 ×
4
c 1,2 ×
4
c 3,4 ×
4
=
@c c 1,3 .
c3 c1
c
c c2
c c4
4 c @
c @ c3,4 4 c
4 @
∼
= 1,2 × 1,2 × @c
@c c3 3
@c
c
@
1
5.4 Tensorial constructions 245
c c2
c c4
c4
1,2 c @ c3,4 ×
4 4
∼
= c 1,2 ×
@c c3
c3
c c1
c
c
c 4 c @ c2
@c @
1,2 c @ c3,4 ×
4
∼
=
@
@c @c @
@ @
c @@c @c
1 3
@c
b
c 4 b
@ b2
c @b
∼
= 1,2 c @ c3,4 4
× b
@c b @ b3
1
c @b
This product has already been calculated in Example 5.1. The result is presented
in Figure 5.2. >
Example A particularly nice application of this method is Thiele’s represen-
tation of a free distributive lattice with four generators as a subdirect product.
The nested line diagram obtained thereby is presented in Figure 1.15.
In general, it is true that FCD(n) is isomorphic to the n-th tensor power of the
three-element lattice [395]. This means that FCD(4) can be obtained as the tensor
product of four three-element chains.
We make use of the fact that a tensor product of two three-element chains can
be rewritten as a 2-product:
c
c
c c c @
c1 2 c2
@
c1 ⊗ c2 ∼ 1 c @c ∼
= 2 = c2 × c1
c c @c
c
@c
@
Thereby we obtain
c c c c
FCD(4) ∼
= c1 ⊗ c2 ⊗ c3 ⊗ c4
c c c c
c c
c1 {1,2} c2 c3 {3,4} c4
∼
= c2 × c1
⊗
c4 × c3
c c
246 5 Constructions of concept lattices
c c c c
c1 c3 4 c1 c4 4 c2 c3 4 c2 c4
∼
= c2 ⊗ c4 × c2 ⊗ c3 × c1 ⊗ c4 × c1 ⊗ c3
c c c c
c1 c3 c c
c 1,3 4 c 4 4 c2 4 c 2 c4
∼
= × × × ⊗ .
c 2,4 c3 c1 c1 c3
c2 c4 c c
The tensor product of two four-element chains has already been calculated in
Example 16; the 4-product of the remaining factors was treated in Example 5.1,
see Figure 5.1. Hence the 4-product presented in Figure 5.12 is isomorphic to the
tensor product of four three-element chains, and thus also to the free completely
distributive lattice with four generators. This is how the diagram in Figure 1.15
has been obtained.
d
d d
@
d @d
@
d @d
1 d
@ @ 3
@d @d
@ @
2 d
@d @ d4
@ @
@d d @d
@ @
@d d @d
4
@ @ ×
d @d @d
@ @
d @d @d
@ @
d @d @d
@ @
1 d
@d @ d3
4 @ @ 2
@d @d
@ @
@d @d
@
@d
@
@d
d
Figure 5.12 The free distributive lattice FCD(4) as a 4-product.
5.5 Notes, references, and trends 247
5.1
Section 5.1 follows [399] and the predecessor of this article, [393].
P -products of lattices are a long-standing subject of one of the authors of this
book, see [387] and [388]. Bartenschlager [25] and Thiele [369] in their work
make ample use of the P -product as a mathematical construction tool. [142] uses
P -products to describe rough sets in the language of Formal Concept Analysis.
5.2
The results of this section are based on the doctoral thesis of S. Gürgens [180], we
have, however, changed the notations. It also contains further-reaching results.
Gürgens states an algorithm which determines whether K is the gluing of two
contexts. Furthermore, she studies the simultaneous gluing of several lattices
or contexts, respectively. Her model were the gluings of Boolean lattices in the
theory of orthomodular lattices, cf. Greechie [176].
5.3
Local doubling was introduced by Day [85], first for intervals and then more
generally. It played an important role in the framework of the examination of free
lattices, see also Day [86], Nation [291] and Day, Nation & Tschantz [88]. Day
even stated a concept-analytic version of the construction of interval doubling.
Our representation mainly follows Geyer [170].
The fact that the bracketings form a lattice was first published by Tamari [367].
Later, Huang and Tamari [195] gave a simpler proof. The concept-analytic inves-
tigation goes back to Geyer [171].
5.4
The set of all order-preserving maps from an ordered set P into a complete
lattice L also forms a complete lattice, when ordered point-wise. This lattice is
denoted by LP . Occasionally, a formula is used which establishes a connection
between this lattice and the lattice 2P of all order-preserving maps of P into the
two-element lattice 2 (for which a context construction was given on Page 48):
LP ∼
= 2P ⊗ L.
Chapter 6
Properties of concept lattices
In this section, we focus on the size of concept lattices. A first result is that
this is not easy to predict: Kuznetsov [245] has shown that it is difficult (“#P -
complete”) to compute the number of formal concepts for given contexts. A. Al-
bano and B. Chornomaz have jointly made an important cause of the size of
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 249
B. Ganter, R Wille, Formal Concept Analysis, https://doi.org/10.1007/978-3-031-63422-2_7
250 6 Properties of concept lattices
with k objects and attributes, defined on Page 46. Its concept lattice is isomor-
phic to the powerset lattice of a k-element set and therefore has 2k elements. In
mathematical literature it is known as the “Boolean lattice with k atoms”. We
abbreviate it to B(k) := B(Nck ).
Definition 90 A formal context K is called Nck -free if it does not contain a sub-
context isomorphic to Nck . A complete lattice L is B(k)-free whenever B(k) does
not order-embed into L. ♦
Figure 6.1 shows a lattice without a B(4)-sublattice, which is not B(4)-free.
Figure 6.1 A lattice containing B(4) as a suborder, but not as a sublattice. From [1].
Proposition 112 A formal context K is Nck -free if and only if B(K) is B(k)-free.
Proof It follows from Proposition 38 on p. 119 that B(k) embeds into B(K)
whenever Nck ≤ K. To prove the converse, let (A1 , A01 ), . . . , (Ak , A0k ) be the atoms
of an B(k) embedded in B(K) and let (D10 , D1 ), . . . , (Dk0 , Dk ) be the coatoms of
the same B(k) in such a way that (Ai , A0i ) ≤ (Dj0 , Dj ) ⇔ i 6= j. For each i we can
choose an object-attribute pair (gi , mi ) with gi ∈ Ai , mi ∈ Di and (gi , mi ) ∈ / I
because of (Ai , A0i ) 6≤ (Di0 , Di ). Since (Ai , A0i ) ≤ (Dj0 , Dj ) holds true for all i 6= j,
it follows that gi I mj for i 6= j. It also follows that the gi are pairwise differ-
ent, as are the mj . That is, the objects gi and attributes mj form a contranominal
subcontext of K.
6.1 Size and breadth 251
Corollary 8 K is Nkc -free if and only if |A| < k for every minimal extent generator A.
The number of subsets of an n-element set which have less than k elements is
X n
f (n, k) := .
i
i<k
According to Corollary 8 an Nck -free formal context (G, M, I) with |G| =: n < ∞
cannot have more than f (n, k) minimal extent generators. And since every finite
extent has at least one minimal generator, (G, M, I) cannot have more that f (n, k)
extents.
Birkhoff [49] defines2 the breadth of a finite lattice to be the smallest integer
b such that any join x1 ∨ · · · ∨ xn (n > b) is the join of a subset of b of the xi .
W X of the lattice can be called an irredundant join representation if
A subset
x 6≤ (X \ {x}) holds for all x ∈ X. The breadth of a finite lattice is then the
maximum size |X| of such an irredundant representation.
These notions fit the topic perfectly:
Corollary 9 The breadth of a finite concept lattice is the largest size of a minimal extent
generator and also the largest size of a minimal intent generator.
It is moreover the largest number b for which B(b), the Boolean lattice with b atoms,
can be order-embedded.
Proof Let X be an irredundant join-representation. Then for each xW ∈ X there
mustWbe some join-irreducible element γgx ≤ x for which γgx 6≤ (X \ {x}),
and {γgx | x W ∈ X} also is irredundant. {gx | x ∈ X}00 is the extent of the
formal concept {γgx | x ∈ X} and {gx | x ∈ X} is a minimal generator of
that extent. This means that a minimal generator of the same size is obtained
1 I.e., a minimal generator wrt. the closure system of concept extents of K in the sense of Defi-
nition 16 on p. 14.
2 In fact, Birkhoff gives the dual definition, but mentions an observation by I. Rose that it is
self-dual.
252 6 Properties of concept lattices
b+1
2b ≤ |B(G, M, I)| ≤ · |G|b .
b!
Albano and Chornomaz also show that for a formal context with n objects
f (n, k) not only places an upper bound on the size of the concept lattice, but also
on the number of pseudo-extents and the number of proper premises (of object
implications). The reasons for this are that proper premises are always minimal
generators and that the canonical basis according to Proposition 28 on p. 98 is
not greater than the number of proper premises.
lent) laws
(D∧ ) x ∧ (y ∨ z) = (x ∧ y) ∨ (x ∧ z)
(D∨ ) x ∨ (y ∧ z) = (x ∨ y) ∧ (x ∨ z)
This law is also equivalent to its dual (DV W ). One direction of the law (DW V ),
namely the inequality
^ _ _ ^
xs,t ≥ xs,ϕ(s) ,
s∈S t∈T ϕ:S→T s∈S
holds in every complete lattice, since, for fixed ϕ, s∈S xs,ϕ(s) is always less than
V
or equal to the left-hand side.
We frequently use a stricter version of the law of complete distributivity, which
can however be derived from the one mentioned above. We allow the set T to
vary with s ∈ S; for this purpose we replace T by a family of sets
{Ts | s ∈ S}.
ϕ∈ ×T
s∈S
s
of the direct product of these sets (which we abbreviate as × Ts). This version
of the law (DV W ) then reads
^ _ _ ^
xs,t = xs,ϕ(s) .
s∈S t∈Ts
ϕ∈ ×T s
s∈S
x 7→ (x] ∩ J(D)
describes an isomorphism of D onto the closure system of all order ideals of (J(D), ≤).
Conversely, for every ordered set (P, ≤) the closure system of all order ideals is a com-
pletely distributive lattice D, in which
J(D) = {(x] | x ∈ P }
is supremum-dense.
x 7→ (x] ∩ J(D)
(g, m) 6∈ I
Assume that the left-hand side is strictly greater than the right-hand side. Then
there exist
\ \ \ \
g∈ ( Bs,t )0 and m ∈ ( As,ϕ(s) )0
s∈S t∈T
ϕ∈ ×T s
s∈S
then n ∈ t∈T Bs,t cannot hold for s ∈ S, because of g ∈ ( t∈T Bs,t )0 . Therefore,
T T
results in a contradiction with (h, m) 6∈ I. Hence, the equation follows from the
conditions specified. In order to be able to use complete distributivity to prove
the inverse direction, we first argue that
^ _
(g 00 , g 0 ) = (ϕ(n)00 , ϕ(n)0 )
ϕ n∈M \g 0
holds for every object g ∈ G, provided the maps ϕ under the -operator are
V
chosen as follows:
ϕ∈ ×
(G \ n0 ).
n∈M \g 0
Thus ϕ runs over all maps that assign to each attribute n which is not incident
with g an object ϕ(n) which is not incident with n. Hence, a possible choice is
ϕ(n) := g for all n, whereby we obtain the direction “≥” of the statement. For
the other direction we note that
W n 6∈ ϕ(n) . Hence,
0
n can still less be contained in
the intent of the supremum n∈M \g0 (ϕ(n)00 , ϕ(n)0 ), i.e., this intent is a subset of
g0 .
We have to show that – provided that the concept lattice is completely dis-
tributive – it is possible, for arbitrary g ∈ G and m ∈ M with (g, m) 6∈ I, to
256 6 Properties of concept lattices
find some h ∈ G and some n ∈ M which satisfy the conditions specified in the
theorem.
With the help of the preliminary considerations and by applying the distribu-
tive law we obtain
^ _ _ ^
(g 00 , g 0 ) = (ϕ(n)00 , ϕ(n)0 ) = (k 00 , k 0 ).
ϕ n∈M \g 0 n∈M \g 0 k∈G\n0
The lattice of the subspaces of a vector space, or more generally the lattice of the
submodules of a module, has a particular structural property: it satisfies the mod-
ular law. The lattices of normal subgroups of groups are also modular. A weaker
x∧y ≺y ⇒x≺x∨y
holds for each two elements x, y. It is modular if it satisfies the following law for
all x, y and z:
x ≤ z ⇒ x ∨ (y ∧ z) = (x ∨ y) ∧ z,
and graded if there is a rank function r(x) assigning a natural number to each
element of L with
x ∧ y ≺ x, y ⇒ x, y ≺ x ∨ y,
x ∨ (y ∧ z) = (x ∨ y) ∧ z.
However, the left-hand side of this equation equals x, the right-hand side equals
z, which is contradictory to x < z.
6.2 Distributivity, modularity and semimodularity 259
g• := {x ∈ G | γx < γg}.
260 6 Properties of concept lattices
Theorem 50 For a doubly founded concept lattice L := B(G, M, I), the following con-
ditions are equivalent:
1. L is semimodular.
2. L satisfies the strong condition of semimodularity.
3. The following exchange condition holds in (G, M, I):
Note: Contrary to the formulation of the theorem, the proof only uses one of
the conditions of foundedness. If we add the other one, we can slightly improve
the result. For instance, in condition (4) (h, p) ∈
/ I can in this case be replaced
by h % p. This makes it possible to show that factor lattices of doubly founded
semimodular lattices are again semimodular.
Proof 1 ⇒ 2: Let y, x and z be elements with x < z and y ∨ x = y ∨ z and
y ∧ x = y ∧ z. First we show that there is an element d with y ∧ x ≺ d ≤ y. Because
of the foundedness thereWexists an element s minimal with respect to s ≤ y,
s 6≤ y ∧ x. s is necessarily -irreducible, and s∗ ≤ y ∧ x, i.e., s ∧ (y ∧ x) = s∗ ≺ s.
If now we set d := s ∨ (y ∧ x), we obtain y ∧ x ≺ d from the first condition of the
proposition. A repeated application of the condition yields x ≺ x ∨ d. x ∨ d ≤ z
would imply z ≥ x ∨ d ≥ d ≥ s and, because of y ≥ s, also y ∧ z = y ∧ x ≥ s,
which would be contradictory to the definition of s. Hence z ∧ (x ∨ d) = x. 2 ⇒
1 has already been proved in Proposition 115.
3 ⇒ 1: Let x, y be concepts with x ∧ y ≺ y and x = (A, B), and let furthermore
z be a concept between x and x ∨ y: x < z < x ∨ y. The foundedness W yields an
element s which is minimal with respect to s ≤ y, s 6≤ x, s then is -irreducible
and thus an object concept, i.e., s = γh for some h ∈ G. Because of s∗ ≤ x we
have h• ⊆ A, furthermore we have x ∨ y = x ∨ s. If we now choose an object g
which is contained in the extent of the concept z but which does not form part
of A, we obtain
g 6∈ (A ∪ h• )00 , g ∈ (A ∪ {h})00 ,
from which by (3) it follows that h ∈ (A∪{g})00 . This extent is however contained
in that of z, which results in a contradiction with γh 6≤ z.
1 ⇒ 3: Trivially, condition (3) is satisfied if g is reducible. Therefore, we can
restrict ourselves to the case (γg)∗ ≺ γg. The preconditions of (3) describe three
6.2 Distributivity, modularity and semimodularity 261
concepts, namely (A00 , A0 ), ((A ∪ {h})00 , (A ∪ {h})0 ) and ((A ∪ {g})00 , (A ∪ {g})0 )
with A00 ⊂ (A ∪ {h})00 ⊆ (A ∪ {g})00 . Since g• ⊆ A is true, γg ∧ (A00 , A0 ) = γg ∗ ,
and thus, because of semimodularity,
r(x) := max{r(y) | x ≺ y} − 1.
If r were not a rank function, then there would have to be elements x ≺ y with
r(x) + 1 6= r(y), and, of all such examples, we could choose one with maximal
y, furthermore, by the definition of r, there would be a further upper neighbor
z of x with r(z) = r(x) + 1. By the condition of semimodularity y ∨ z would be
262 6 Properties of concept lattices
In the atomistic case we have g• = ∅ for all g ∈ G. Thus, the exchange condition
simplifies to yield the known form
(SD∨ ) x ∨ y = x ∨ z ⇒ x ∨ y = x ∨ (y ∧ z) = x ∨ z
(SD∧ ) x ∧ y = x ∧ z ⇒ x ∧ y = x ∧ (y ∨ z) = x ∧ z.
In a concept lattice the extremal points of a concept (A, A0 ) are precisely the
object concepts γg with g ∈ A but
g 6∈ (A \ {h | g 0 = h0 })00 .
g 6∈ (A \ {h | g 0 ⊇ h0 })00 .
Theorem 51 For a doubly founded concept lattice L = B(G, M, I), the following state-
ments are equivalent:
1. L satisfies (SD∨ ).
2. For all g, h ∈ G and all m ∈ M we have:
. m and h %
g% . m imply g 0 = h0 .
µm ∨ γg = µm∗ = µm ∨ γh
Now γg ∗ as well as γh∗ is less than or equal to µm. The above equation can only
be true if neither γg ∧ γh ≤ γg ∗ nor ≤ γh∗ . This forces γg = γh.
Now we conversely assume that (G, M, I) satisfies the above-stated condition
for the arrow relations and show that B(G, M, I) is join-semidistributive. Let
furthermore x, y and z be elements of B(G, M, I) with
x ∨ y = x ∨ z > x ∨ (y ∧ z).
t ≥ x ∨ (y ∧ z), t 6≥ x ∨ y.
u ∨ ū = a = u ∨ x
u ∨ (ū ∧ x) = a,
we have
h0 ⊆ g 0 ⇔ γh ≥ γg ⇔ γh ≥ ū ⇔ γh 6≤ u.
Hence, in this case, A \ {h | g 0 ⊇ h0 } is entirely contained in the extent of u, from
which follows the desired base point property:
g 6∈ (A \ {h | g 0 ⊇ h0 })00 .
5. L is meet-distributive.
6. Every element has a unique irredundant -representation.
W
ϕ : L → P(X)
a ≤ t, b 6≤ t, b ≤ t∗ ,
a ≤ b ∧ t ≺ b.
vx is -irreducible (or the smallest element of L), since there can only be one
W
lower neighbor u of v because the image
e
@
@
@
g % m, g . n ⇒ g .
%m semi convex @ g . m, h % m ⇒ g .
%m
e e e
@ @ @
dually
@ SD∧ @ semimodular
@ @ @
semimodular SD∨ @
e e @ e @e @e
HH @ join-
H
meet-
@ @ @
H @ distributive
distributive
@ @ @ HH
@ @ @ H@
e @e @e @e @e
b
H H
b @
b @b @
@
@ @ b b@ b
b @b
@
b@ b b @
b @
@ @ @
b
B b@e b
@b e
@ @ @
b A@b @ modular @
b b b
@ A@
b A@ @ b @ @@ b
b b b AA@ b@ b
@ A @
b A@b @ b@A b @ A @ @b @ b A@ @ b@ A
@
@b b @ b b
@ b@A b
A @
@ A
b b@ A b @ @
@ b @A b
AA b @
b @ @
@@
b b @ A
A
@b @b @b
@
@b @ A
@A b b @b
distributive@A e @b @ @b
@ @ @b
@@b
Figure 6.2 Generalizations of the Distributive Law.
Figure 6.2 shows the implications between the above mentioned lattice prop-
erties (for doubly founded complete lattices). Two of the attributes still lack an
explanation: According to A. Day, a lattice is semiconvex if it satisfies the follow-
ing condition:
B stands for the property of being a bounded homomorphic image of a free lat-
tice. This property has a simple characterization in the language of contexts: it
is equivalent to the fact that the objects and attributes of the (reduced) context
can be ordered in such a way that every % . is on the diagonal, every % below the
diagonal and every . above the diagonal.
The convex-ordinal scales C(P,≤) (see Section 1.4), at least if (P, ≤) is finite,
are doubly founded. The extents are the convex subsets of (P, ≤), the extremal
points are the maximal and minimal elements of such subsets. Hence Condition
6.3 Semidistributivity and local distributivity 269
(3) of Theorem 52 is satisfied and therefore finite convex-ordinal scales are meet-
distributive.
d
1
d
##
#d
d#
# cc
#
#d
d d## c d
cc
#
#J##Jc c ##J
# d#J d2Jcdc #d
cc Jd
d #J
# c # cc d #J #
#
# # c J# # #
dc # # Jd
c
#J Jc
# #
d# J d# # d# c J d## 3
c c c
d
c
#c# c## #
c
J c
# c # c c J ###
J d#
c # J #
cJ d#
4 cc d#
c ccJ d## 6
7 5 #
cJ #
ccJ d#
The Anti-exchange Axiom holds in the closure system of the convex sets of an
arbitrary metric space. But it also holds in other connections, for example, if we
can assign a weight wt(g) to every object g, such that wt(g) = wt(h) only if g 0 = h0
and such that
g 6∈ A00 , g ∈ (A ∪ {h})00 ⇒ wt(g) ≥ wt(h).
The last condition can be interpreted as saying that the weight of g must be at
least as big as that of h if g is generated by means of h. A simple example in
this connection is presented in Figure 6.3. If G ⊆ N is an arbitrary set, then a
subset T ⊆ G will be called additively saturated if from a, b ∈ T, a + b ∈ G
it already follows that a + b ∈ T . The additively saturated subsets of G form a
closure system which with wt(g) := g obviously satisfies the above-mentioned
condition. Hence the corresponding lattice is meet-distributive. Figure 6.3 shows
an example with G := {1, 2, . . . , 7}.
270 6 Properties of concept lattices
6.4 Dimension
Order dimension
dim(P, ≤) = n
As a direct corollary of Proposition 39 we obtain:
Proposition 116 There is an order embedding of B(G, M, I) in a product
×(L , ≤)
t∈T
t
if and only if there are pairs of maps (αt , βt ), t ∈ T , with the following properties:
1. αt : G → (Lt , ≤), βt : M → (Lt , ≤),
2. (g, m) ∈ I ⇒ αt g ≤ βt m for all t ∈ T ,
3. (g, m) ∈
6 I ⇒ αt g 6≤ βt m for some t ∈ T .
We can express this fact differently by replacing the maps αt and βt by rela-
tions Jt with (g, m) ∈ Jt : ⇔ αt g ≤ βt m:
4 Meaning an order embedding according to Definition 6.
6.4 Dimension 271
×(L , ≤)
t∈T
t
Ferrers dimension
Nevertheless, in the case of small contexts it can be carried out by hand. In this
connection it is convenient to make use of the fact that the complement of a Fer-
rers relation is again a Ferrers relation. Hence the Ferrers dimension of (G, M, I)
is also equal to the smallest number of Ferrers S relations Ft covering the empty
cells of the cross table, i.e., with G × M \ I = t∈T Ft . It is, however, not always
possible to choose this covering to be disjoint.
Theorem 54 The Ferrers dimension of (G, M, I) is equal to the order dimension of the
concept lattice B(G, M, I):
The order dimension of an ordered set (P, ≤) is equal to the Ferrers dimension of
(P, P, ≤):
dim(P, ≤) = fdim(P, P, ≤).
Proof This follows immediately from the Propositions 117 and 118, because, ob-
viously, a complete lattice can be embedded in a chain if and only if it is a chain
itself.
× × 1 × 1 1 1 (3, 0, 0, 0) (3, 0, 3, 3)
2 × × 4 × 1 1 (2, 1, 0, 1) (3, 3, 0, 3)
2 3 × × 2 × 1 (1, 2, 1, 0) (2, 2, 1, 3)
2 3 3 × × 3 × (0, 2, 3, 0) (3, 2, 3, 0)
× 3 3 4 × × 1 (1, 0, 2, 1) (2, 1, 3, 2)
2 × 2 2 2 × × (0, 3, 0, 0) (1, 3, 2, 2)
× 3 × 4 4 4 × (0, 0, 1, 3) (0, 3, 3, 3)
Figure 6.4 Point-line context of the projective plane PG(2,2), with Ferrers relations. The
Ferrers dimension, and thus the order dimension of the plane, is 4; a covering of the 28
empty boxes with less than four Ferrers relations is impossible, since each Ferrers relation
in this example can have eight elements at most. On the right is an embedding of PG(2,2)
into a product of four chains; the first column gives the images of the points, the second
those of the lines.
k-dimension
× × 1 × 1 1 1
2 × × 2 × 1 1
3 3 × × 3 × 3
3 3 4 × × 5 ×
× 4 4 4 × × 4
2 × 2 2 2 × ×
× 5 × 5 5 5 ×
Figure 6.5 The 3-dimension of PG(2,2) is 5. In arithmetic terms a covering with four Fer-
rers relations of the length 3 would be conceivable, however, such a covering does not
exist. The 2-dimension is 7.
Set representations
g I m ⇔ αg ∩ βm 6= ∅.
g I m ⇔ αg ∩ βm = ∅
is satisfied. ♦
In Section 1.4 we had defined the contexts for free distributive lattices by
means of set representations.
Ft := {g | t ∈ αg} × {m | t ∈ βm}
= {(g, m) | t ∈ αg ∩ βm}
⊆ G×M \I
We know already from the observations following Proposition 117 that in or-
der to determine the Ferrers dimension of a doubly founded context B(G, M, I),
it suffices to cover those pairs (g, m) of the complement of I for which g . m or
g % m holds. In fact, we can even restrict ourselves to the double arrows, since
they play the role for the Ferrers dimension which the critical pairs known from
order theory play for the order dimension.
6.4 Dimension 275
fdimk (G, M, I) ≤ n
also holds.
Proof First we describe the possibility of suitably extending a given Ferrers re-
lation F . The basic idea is that the Ferrers condition is not affected if we double
some row or column of the context. Formally, this can be described as follows:
For a Ferrers relation F ⊆ G × M and objects g, h ∈ G,
F ∪ {(g, m) | (h, m) ∈ F }
F ∪ {(g, m) | (g, n) ∈ F }.
we had earlier introduced the term preconcept of the context (G, M, I). Hence the
set dimension is equal to the smallest number of preconcepts (more precisely:
sets A × B, where (A, B) is a preconcept) whose union fills up I. Since, however,
every preconcept can be extended to a concept, the set dimension is also equal
to the smallest number of concepts whose union is I. We give an example:
Example We want to find out whether the digits of the seven-segment display
a b c d e f g
× ×⊗×××
⊗×
a ××× ⊗×
d f ××⊗ ××
b ⊗ × ××
e g ×××× ×
c ××××× ×
⊗ ××
×××××××
×××× ××
Figure 6.6 The seven-segment display and the context of the digits. The crosses sur-
rounded by circles mark a blocking set.
For this purpose we consider the context in Figure 6.6, whose objects are those
digits and whose attributes are the segments of the display; the incidence is ex-
plained in the obvious way. The intents naturally correspond to partial figures of
the display; a set of concepts fills out I if and only if every object intent can be rep-
resented as the union of (concept) intents from this set. Thus, the task of getting
along with as few partial figures as possible is equivalent to the determination
of the set dimension. In fact, six intents are sufficient, namely
The fact that a smaller number cannot suffice is evidenced by the “blocking
set” of incidences marked in Figure 6.6: No two of these crosses can belong to
a common preconcept. >
6.4 Dimension 277
where the set F of “factors” is initially unknown. The aim is usually to get by
with as few factors as possible.
A look at Definition 33 shows that for each factor f ∈ F the set of pairs
f IGF × f IF M
g I m ⇔ ∃n∈δ(m) g J n.
a b c d e f g
× × × × × × × × × × × a b c d e f g
× × ×
× ×
× × × × × × × × ×
× × × × × × ×
× × × ×
× ×
× × × × = × × × ·
× × × × × × × × ×
× × × × × × × × × × × ×
× × × × × × ×
× × × × × × × × × × × × ×
× × × × × × × × × × ×
Figure 6.7 The formal context from Figure 6.6 allows a Boolean matrix factorization with
six factor attributes.
W
-dimension
Embeddings in direct productsWof chains are also of interest for W the drawing of lat-
tice diagrams. In this context, -embeddings, i.e., injective, -preserving maps,
have proved helpful, and we are faced with the question of the existence of such
embeddings. The theoretical background of this question is elaborated in Chap-
ter 7 in a different context. We give the following result without proof, because
it can be obtained as a special case of the Propositions 132 and 134.
Theorem 58 There exists a -embedding of a finite lattice L in the direct product of
W
n chains of the lengths l1 , . . . , ln if and only if M (L) can be covered by n chains of
cardinality m1 , . . . , mn , where mi ≤ li holds for i ∈ {1, . . . , n}.
As an immediate consequence, we can W determine two parameters for this
problem of embedding: If we define the -dimension to W be the smallest number
of chains in whose product the lattice can be embedded -preservingly and the
-rank to be the least length of such a product, the theorem gives us satisfactory
W
information on those numbers, at least for finite lattices. For this purpose we use
a well known result of Dilworth ([81], p. 3), which says that the smallest num-
ber of chains by which a finite ordered set can be covered is equal to the width
of this ordered set.
Corollary 11 InW the case of a finite lattice L, the -dimension is equal to the width of
W
M (L) and the -rank is equal to the cardinality of M (L).
6.5 Notes, references, and trends 279
6.1
6.2
the distributive law coincide in the doubly founded case. A connection is estab-
lished between Theorem 47 and topological statements. This theorem was taken
from [396] and goes back to ideas of Raney [310]. Finite distributive concept lat-
tices are described in [395]. That distributivity can be described by means of the
arrow relations has been known for quite a long time in the version of Theorem
48(9). The elegant concise version in Theorem 48(3,4) also goes back to Erné.
There is a book by Stern [352] on semimodular lattices. The characterization of
semimodularity by means of the arrow relations was derived by Skorsky from
a result of Faigle and Herrmann [121]. For the arrow relations in contexts of
modular lattices, see Doerfel [99].
6.3
6.4
A standard reference for the dimension theory of ordered sets is the book by
Trotter [370]. The fact that the order dimension of (P, ≤) is equal to that of the
Dedekind-MacNeille-completion has been known for quite a long time, compare
[20], [337]. Baldy & Mitas [19] have generalized this.
The terms Ferrers relation and -dimension go back to Riguet [320] and Cogis
[78], [79], The works of the latter already contain parts of the content of Theo-
rem 54. Bouchet [65] proves the theorem for ordered sets, the generalized version
can be found in [403]. With regard to the Ferrers dimension see also Doignon,
Ducamp & Falmagne [100] and Koppen [226]. Applications can be found in
Reuter [315], [316], Ganter, Nevermann, Reuter and Stahl [153]. Closely related
b b relations are the interval orders. These are ordered sets (P, ≤) in
to the Ferrers
which b b cannot be embedded. Formally, the condition reads:
The mappings between concept lattices which are important for structure com-
parison are primarily the complete homomorphisms. In Section 3.2 we have
worked out the connection between compatible subcontexts and complete con-
gruences, i.e., the kernels of complete homomorphisms. Another approach is to
connect morphisms between lattices with morphisms between formal contexts.
But what are suitable morphisms between formal contexts? For isomorphisms
and thus especially for automorphisms the answer is obvious, even if mapping
pairs are required. We show in the first section how to compress concept lattice
diagrams if there are non-trivial automorphisms. In a folded diagram, the nodes
do not represent the individual concepts, but entire orbits of the automorphism
group.
In the second section we then ask for suitable general morphism terms for
formal contexts. We start with structure-preserving or -reflecting mappings and
come across the extensionally continuous mappings, which become important in
the third section under the name scale measures and even fill the fourth section.
Mapping pairs are also considered again. However, the most natural morphisms
for formal contexts are probably the bonds that have already been considered in
section 3.4.
Scale measures are needed for conceptual measurement, i.e. for the process of
mapping an observation (noted as a formal context) into a scale (which is also
a formal context). The third section summarizes the most important properties,
the fourth contains first approaches to a conceptual measurement theory.
g I m ⇔ α(g) J β(m).
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 283
B. Ganter, R Wille, Formal Concept Analysis, https://doi.org/10.1007/978-3-031-63422-2_8
284 7 Context comparison and conceptual measurement
induces a lattice isomorphism of B(K1 ) onto B(K2 ). If both contexts are reduced,
then every lattice isomorphism is induced by one (and only one) context iso-
morphism. More generally: If the contexts are clarified, a lattice isomorphism ϕ
is induced by a context isomorphism if and only if ϕ surjectively maps object
concepts onto object concepts and attribute concepts onto attribute concepts.
An observation by W. Xia shows that in turn we can interpret the isomor-
phisms themselves as concepts of a suitable context.
Definition 98 For contexts K1 := (G, M, I) and K2 := (H, N, J) we define
∼
K1 × K2 := (G × H, M × N, ∼)
with
(g, h) ∼ (m, n) :⇔ (g I m ⇔ h J n). ♦
as well as
B ∼ = {(g, h) | g I m ⇔ h J n for all (m, n) ∈ B}.
Hence, a pair (α, β) of bijective maps is an isomorphism of K1 onto K2 if and only
if α ⊆ β ∼ or, equivalently, if β ⊆ α∼ . This, however, also implies that α = β ∼
and β = α∼ , since, if (g, h) ∈ β ∼ , then
hJβm ⇔ g I m ⇔ αg J βm
for all m ∈ M , from which we infer that hJ = (αg)J . If K2 is clarified, this implies
h = αg and we obtain α = β ∼ .
Furthermore, we learn from the theorem that, in the case of clarified contexts,
an isomorphism (α, β) is determined already by each of its two components.
A further application of Theorem 6 on p. 82 allows us to calculate a concept
lattice “modulo automorphisms”. This means: If Γ ≤ Aut(G, M, I) is a group
7.1 Automorphisms and foldings 285
which, because of
A = (A ∩ {1, . . . , i − 1})00
would imply that A ⊆ α(A) and thus A = α(A), contradictory to A < α(A).
Hence, j < i must hold. But then we have
Lattice foldings
Example Figure 7.1 shows the lattice of the subgroups of the alternating group
A5 . However, this is not a usual diagram of the subgroup lattice, because firstly,
only nine elements are shown, although the group A5 has 59 subgroups. Sec-
ondly, it is not a lattice line diagram as introduced in Section 0.1, because the
elements labeled Z2 and Z3 have no supremum. And finally, one of the edges
(far left) bears an unusual annotation “(24)”. Further explanation is obviously
needed to make the diagram comprehensible.
Represented are the nine orbits of the automorphism group Γ of A5 , which is
isomorphic to S5 . Each circle of the diagram represents several subgroups that
286 7 Context comparison and conceptual measurement
eA5
HH
HH
HH
H
D3 eH eA4
HH e
D5
HH
(24)
HH eV4
H
HH
Z3 e H e eZ5
HH Z2
HH
H
HH
HH e
S1
Figure 7.1 The lattice of subgroups of the alternating group A5 “modulo automor-
phisms”. The representatives of the subgroup orbits are: S1 := {id}, Z2 := {id, (01)(23)},
Z3 := {id, (012), (021)}, V4 := {id, (01)(23), (03)(12), (02)(13)}, Z5 := h(01234)i,
D3 := h(01)(23), (014)i, D5 := h(01)(23), (01234)i, A4 := h(01)(23), (012)i, A5 :=
h(012), (01234)i.
can be mapped to each other by the automorphism group. The diagram shows
a factor structure that is completely different from the homomorphic images dis-
cussed in section 3.1. A representative for each subgroup orbit is given in the
figure caption.
The order relation is defined as follows: an orbit is less or equal to another
one, if one of its elements is a subgroup of some element of the second orbit. It
is easy to see that this implies that every element of the first orbit is contained in
some element of the second. And it also implies, that every representative of the
second orbit contains some element of the first. The order relation is independent
of the choice of representatives.
The additional annotation is needed to express which elements of one orbit
are contained in which elements of the second. It is surprising that in Figure 7.1 a
single edge label is sufficient. This is only possible because the labeling has been
strictly simplified, and does depend on the choice of representatives.
In Figure 7.1, the representatives were chosen in such a way that they almost
always actually reflect the order relation, i.e. one orbit is contained in another if
the chosen representative is a subgroup of the second representative. The excep-
tion is Z3 , because this group is not a subgroup of the group D3 , which represents
the orbit above it. But D3 does contain the group (24)Z3 (24), which is obtained
from Z3 by conjugation with the permutation (24). This is the reason why the
element (24) is entered at the corresponding edge of the line diagram. >
In the case of lattices with many automorphisms a diagram of the orbits is
often easier to read than a diagram showing all elements of the lattice. Such di-
7.1 Automorphisms and foldings 287
agrams “modulo automorphisms” are used in group theory, but can be applied
more generally.
Definition 99 Let (P, ≤) be a finite ordered set and let Γ ≤ Aut(P, ≤) be a sub-
group of the automorphism group of (P, ≤). A folding of (P, ≤) under Γ , also
called a Γ -folding of (P, ≤), is
where
– T is some transversal of the orbits of Γ on P ,
– a ≤f b : ⇔ ∃γ∈Γ a ≤ γ(b) (for a, b ∈ T ),
– (G, ◦) with G := {gγ | γ ∈ Γ } is a group for which γ 7→ gγ is an isomorphism
from Γ to (G, ◦), and
– λ : T × T → P(G) is defined by
S1 ◦ g ◦ S2 = {s1 ◦ g ◦ s2 | s1 ∈ S1 , s2 ∈ S2 }
The fact that the double cosets partition the group allows for a much more read-
able annotation.
Definition 100 A simplified annotation λ• corresponding to λ specifies
– for each element a ∈ P the group Ga = λ(a, a), usually by annotating a struc-
ture whose automorphism group is Ga , and
– for every pair a < b a transversal of the double cosets in
[
λ(a, b) \ λ(a, c) ◦ λ(c, b).
a<c<b
♦
Note that λ• (a, b) may be empty. The convention is that such pairs are not
listed at all. Likewise, if only the neutral element occurs as a representative for
neighboring pairs a ≺ b, nothing is written. It also happens that λ• (a, b) 6= ∅ even
if a 6= b and a is not a lower neighbor of b. It may then be necessary to add an
edge to the folding diagram that carries this annotation. Figure 7.3 shows such
a situation.
(243)
(243)
(23)
2 1
3 4
Figure 7.2 An S4 -folding of the lattice of all graphs on four points, with simplified anno-
tation.
Example Figure 7.2 shows the powerset lattice of a six-element set, not as a usual
lattice diagram, but as a lattice folding. Undirected graphs on the four-element
7.1 Automorphisms and foldings 289
set {1, 2, 3, 4} are shown, i.e. subgraphs of the complete graph K4 . This graph has
six edges, and each of the 64 subsets of this edge set determines a graph on the
same nodes. However, the folding diagram does not show all 64 graphs, but only
the 11 isomorphism classes, which are the orbits of the automorphism group S4
of K4 , acting on its subgraphs.
At this point, we will explain in detail which groups we are talking about.
The lattice which was folded is the lattice of all 64 graphs on {1, 2, 3, 4}. The
automorphism group of this lattice has a subgroup Γ , which is induced by the
permutations of the node set and which maps graphs to isomorphic graphs. For
the folded structure, the information on how Γ operates on the original lattice is
no longer given, but only that this group (G, ◦) is isomorphic to S4 .
From each graph isomorphism class, a representative was chosen and entered
with its graph diagram, thereby specifying the groups λ(a, a) as being the auto-
morphism groups of these representatives.1
We explain the edge labeling using two examples. The circle labeled is
lower neighbor of the circle labeled , and the edge between these two circles
carries no annotation, meaning that (in the unsimplified annotation λ) this edge
is annotated by the double coset containing the neutral element, i.e. by the prod-
uct of the two automorphism groups. Indeed, is a subgraph of , and thus
the identity is one of the exactly eight embeddings. is not a subgraph of its
upper neighbor , but can be embedded. One such embedding is the annotated
mapping (23), the others fill the double coset of which (23) is a representative.>
Figure 7.3 shows a further simple example: on the left, a lattice whose automor-
phism group is apparently isomorphic to the cyclic group Z3 and on the right,
the orbit diagram with the necessary information.
Now that we are able to fold ordered sets we also would like to unfold them
in a way that reconstructs the original order. Before doing so, we reconstruct the
full annotation.
Proposition 123
[
λ(a, b) = λ(a, a) ◦ λ• (a, c) ◦ λ(c, c) ◦ λ• (c, b) ◦ λ(b, b).
a≤c≤b
1 To avoid the term “isomorphism class”, some also speak of the unlabeled case. In this sense,
Figure 7.2 shows a diagram of the unlabeled graphs on four vertices. However, this may be
confusing, as it is a diagram in which each node is labeled with a labeled graph.
290 7 Context comparison and conceptual measurement
c cZ3
@
c c @c c{id}
@
c c c (012) c{id}
c c c c{id}
@
@c cZ3
@
Figure 7.3 The diagram of orbits (on the right) has an additional edge.
But since the chain length for both a < c and c < b are shorter, these sets are
contained in the RHS of the proposition.
Definition 101 The unfolding of an order folding (T, ≤f , λ, (G, ◦)) is
˙
[
where g ◦ Ga ≤r h ◦ Gb : ⇔ g −1 ◦ h ∈ λ(a, b).
G/Ga , ≤r ,
a∈T
♦
Remark: For any a ∈ T , G/Ga equals the set {g ◦ Ga | g ∈ G} of all cosets of
the “stabilizer” subgroup Ga . The intuition behind this is that a represents one
isomorphism type and the cosets g ◦ Ga stand for all possible isomorphic copies
of this type. However, Ga and Gb can be the same, even if a and b are different.
For example, in Figure 7.2 the representatives and have the same auto-
morphisms. This is why a disjoint union is required in the above definition. But
the somewhat vague definition of a disjoint union can lead to difficulties in no-
tation, which can be overcome by using the “dot notation” as in Definition 29 on
p. 41, replacing ˙ p∈P G/Gp by p∈P ({p} × G/Gp ). Having said this, we will
S S
avoid this exact but more cumbersome notation and write as if cosets g ◦ Ga and
h ◦ Gb are always distinct when a 6= b.
Theorem 61 The unfolding of any folding of a finite ordered set (P, ≤) is isomorphic to
(P, ≤).
Proof Let Γ ≤ Aut(P, ≤) be a subgroup of the automorphism group, Y be a
transversal of the Γ -orbits and (T, ≤f , λ, (G, ◦)) be a folding of (P, ≤) under Γ ,
as in Definition 99. Recall that γ 7→ gγ denotes a group isomorphism from Γ to
(G, ◦). We have to show that
˙
[
G/Ga , ≤r
a∈T
7.1 Automorphisms and foldings 291
××
× ×
××
××
× ×
×
×
××
×
××
×
× ×
×
×
(bc)
×× ×× ×× ×
× × × × ×
× × ×
(bc)
××
×
×× ×
× × @ ×
(ac)@
××
×
×× ×
× ××
×× ×
a b c
Figure 7.4 An S3 -folding of the lattice of the 61 closure systems on the set {a, b, c}. Each
closure system is the system of intents of an object-reduced formal context, which is
unique up to permutations of the attribute set {a, b, c}. The context at the least element
has an empty object set. From [178], modified. An unfolded diagram can be found at
Higuchi [191].
is an order isomorphism. It is
gγ ◦ Ga ≤r gδ ◦ Gb ⇔ gγ−1 ◦ gδ ∈ λ(a, b)
⇔ gγ −1 δ ∈ λ(a, b)
⇔ a ≤ γ −1 δ(b)
⇔ γ(a) ≤ δ(b),
7.2 Morphisms
g I m ⇒ α(g) J β(m)
for all g ∈ G, m ∈ M ,
- incidence-reflecting if
g I m ⇐ α(g) J β(m)
for all g ∈ G, m ∈ M ,
- continuous if α is extensionally continuous and β is intensionally continuous,
- concept-preserving if, for every concept (A, B) ∈ B(K1 ), the pair
(β(B)0 , α(A)0 )
is a concept of K2 , and
- concept faithful if it is both concept-preserving and continuous.
♦
Even if (α, β) is incidence-preserving and -reflecting, the two maps need not
be injective. However, in this case from α(g) = α(h) follows that g 0 = h0 and
7.2 Morphisms 293
dually; i.e., this kind of map is “injective up to clarification”. If both maps are in-
jective, incidence-preserving and -reflecting, then (α, β) may be called a context-
embedding. It has already been explained in Section 3.1 that this results in
order-embeddings, and that order-preserving maps correspond to incidence-
preserving maps. The following proposition follows directly from Proposition 38
on p. 119 and the text that follows it.
Proposition 124 If (α, β) : K1 → K2 is incidence-preserving, then the two maps
_ ^
ϕ(A, B) := γαg and ψ(A, B) := µβm
g∈A m∈B
both are order-preserving. If (α, β) is also incidence-reflecting, then these mappings are
order-embeddings.
We cannot formulate the opposite direction so smoothly. It is obviously not
the case that every order embedding also entailes a context-embedding, as even
the simplest examples show.2 But one can also specify the reason for this: Order-
preserving mappings (and thus also order embeddings) depend only on the
isomorphism type of the embedding order, not on its representation as a con-
cept lattice. If you allow the second concept lattice, i.e. B(K2 ), to be replaced by
an isomorphic concept lattice, then you can also represent any order-preserving
mapping by an incidence-preserving one. The formal context replacing K2 can
always be chosen as the largest possible one, which is the ordinal scale of B(K2 ).
This is the statement of Proposition 38 and the preceding text on page 119, which
gives the following:
Proposition 125 If ϕ : B(K1 ) → B(K2 ) is order-preserving. then α := ϕ ◦ γ and
β := ϕ ◦ µ are maps for which
is concept faithful if and only if the following condition is satisfied for all h ∈ H, n ∈ N :
ϕ : B(G, M, I) → B(H, N, J)
S := Rϕ = {(h, m) | γh ≤ ϕµm} ⊆ H × M
ARJ = B S , B SJ = AR
ϕ(A, B) := (B S , AR )
M N
G I R := Rϕ
H S := Rϕ J
α : G → P(N ), β : M → P(H)
to each homomorphism
ϕ : B(G, M, I) → B(H, N, J)
through
αg := g R = {n | ϕγg ≤ µn}
βm := mS = {h | γh ≤ ϕµm}.
It is not difficult to characterize the pairs of maps which result in this way from
complete homomorphisms by means of the conditions stated in Theorem 63.
The symmetrical situation in Theorem 63 permits further variations. We can
also describe the bonds through maps in the other direction, i.e., α : H → P(M ),
β : N → P(G), and so on. We shall give only one further example of this kind:
296 7 Context comparison and conceptual measurement
ϕ : B(G, M, I) → B(H, N, J)
the definition
g ∈ αh : ⇔ γh ≤ ϕγg, m ∈ βn : ⇔ ϕµm ≤ µn
Infomorphisms
These morphisms were introduced by Barwise and Seligman [28]. They are the
same as the Chu maps, see H. Mori [288]. Replacing the second context by its
dual gives the context Galois connections of W. Xia [425].
Definition 103 An infomorphism from a formal context K1 := (G, M, I) to
K2 := (H, N, J) is a pair (~f, f~) of mappings ~f : G → H, f~ : N → M, such
that
g I f~n ⇔ ~fg J n
holds for all g ∈ G, n ∈ N . ♦
Proof The defining condition in Definition 103 retains its truth value when I and
J are replaced by r
I and J
r. For an object n of Kd and an attribute g of Kd we find
2 1
that
n J d ~fg ⇔ f~n I d g
7.2 Morphisms 297
Relational morphisms
ϕ(x) ≤ y ⇔ x ≤ ψ(y).
It is also shown in this proposition that the two mappings ϕ and ψ determine
each other and that every -morphism is part of an adjunction, as is every -
W V
morphism.
We have also discussed other relations between formal contexts and resulting
mappings between concept lattices, such as dual bonds and relational Galois
connections. From the point of view of category theory, however, adjunctions
are particularly suitable as morphisms, because they can be concatenated, as can
bonds.
This results in two equivalent categories, namely the category of formal con-
texts with the bonds as morphisms and the category of complete lattices with
adjunctions. A careful presentation of the equivalence can be found in Borg-
wardt [64] together with details about special properties of these categories.
298 7 Context comparison and conceptual measurement
g Iσ m :⇔ σ(g) IS m.
Now, the definition says that σ is a S-measure if and only if every extent of Kσ is
also an extent of K; σ is full if and only if K and Kσ have the same extents. Since
the context Kσ is defined on the same object set as K, we can imagine the two
contexts joined together to form the apposition
K Kσ ,
whose extents are the same as those of K, provided that σ is a measure. In this
way a S-measure is understood as the possibility of extending the given context
by attributes from the scale S without changing the extents. σ is full if the new
attributes render the old ones dispensable.
Proof (1) ⇒ (2): Every scale measure satisfies condition (2), since, for every
A ⊆ G, σ −1 (σ(A)00 ) is an extent containing A and thus also A00 .
(2) ⇒ (3): A → B ⇔ B ⊆ A00 ⇒ σ(B) ⊆ σ(A00 ) ⊆ σ(A)00 ⇒ σ(A) → σ(B).
(3) ⇒ (1): If U is an extent of S and g is an arbitrary object from (σ −1 (U )), i.e.,
with σ −1 (U ) → g, this yields U → σg, i.e. σg ∈ U and consequently g ∈ σ −1 (U ),
hence this set must be an extent of K.
σ is full if and only if for A ⊆ G it always holds that A00 = σ −1 (σ(A)00 ); this is
however equivalent to
g ∈ A00 ⇔ σ(g) ∈ σ(A)00 ,
i.e., to
A → g ⇔ σ(A) → σg.
7.3 Scale measures 299
We do not really have to check the definition of the S-measure for all extents. It
suffices that the preimages of the attribute extents of S are extents of K (since the
preimage of an intersection of sets is the intersection of the preimages). Likewise,
an S-measure is already full if every column extent of K is the preimage of an
extent of S. If we call a subset T of the attribute set of a context K dense in the
case that the set {µm | m ∈ T } is infimum-dense in B(K), then we can continue
as follows: A scale measure is full if and only if the set
is dense.
A surjective S-measure is not automatically full. Indeed, every scale measure
σ can be replaced by a surjective one if we switch to a subscale (by a subscale of
a scale S we understand a subcontext (T, MS , IS ∩ (T × MS )) with T ⊆ GS ). This
is the content of the following proposition:
Proof The extents of the subscale are precisely the sets of the form U ∩ σ(G),
with U being an extent of S.
From a scale measure σ we obtain two maps between the concept lattices of
the corresponding contexts K and S, i.e., one in each direction. This is described
by the following two propositions. From the circumstance that the set mapping
σ −1 is intersection-preserving we immediately infer:
defines a -morphism
W
σ̃ : B(K) → B(S).
σ̃ maps the object concepts of K onto object concepts of S. If S is a scale
W in which g 6= h
always implies g 0 6= h0 (for all g, h ∈ GS ), then, conversely, every -preserving map of
300 7 Context comparison and conceptual measurement
B(K) to B(S) with this property results from an S-measure in the manner specified. σ
and σ̃ uniquely determine each other. σ is full if and only if σ̃ is injective.
σ(x) = g :⇔ ψ(x00 , x0 ) = (g 00 , g 0 ).
Hence, every map of this kind is induced by a measure. The remaining state-
ments follow from Proposition 11 on p. 19.
Definition 105 If S1 and S2 are scales with the same scale values, i.e., with GS1 =
GS2 , we call S1 finer than S2 if every extent of S2 is also an extent of S1 . S2 then is
coarser than S1 . S1 and S2 are called (scale-)equivalent if there is a full bijective
S2 -measure of S1 .3 ♦
If σ is a bijective full measure, so is σ −1 . Hence, the equivalence of scales
is symmetrical. Since the concatenation of (full) scale measures again yields a
In Section 1.4 we have suggested the interordinal scale for the ≤2≥6
scaling of such attributes, however, it is usual first to scale
1 ×
coarser, for example with the threshold scale (displayed on the
2 ×
right), which only uses two elements of the interordinal scale.
3
In this way we obtain an approximate impression of the results.
However, this coarsening is quite correct: According to Propo- 4
sition 132,Wthe concept lattice obtained in this way is an image 5
(under a -morphism) of the concept lattice which is the result 6 ×
of interordinal scaling. 7 ×
The scales Nn , On and Ma,n−a are also (equivalent to) coarsenings of the in-
terordinal scale In . The finest scale with n values is the contranominal scale Ncn .
In order to describe the role of the scale measures in the case of plain scaling,
we first convince ourselves that the semiproduct
×
` [
a
Sj := Gj , Ṁj , ∇
j∈J j∈J j∈J
with
(gj )j∈J ∇ (k, m) : ⇔ gk Ik m
of scales introduced in Definition 31 on p. 43 is also a product in the sense of
category theory, namely in the category of the scales with the scale measures as
morphisms. For this purpose we check that the projections
πk : ×G
j∈J
j → Gk
302 7 Context comparison and conceptual measurement
product map is a scale. The property claimed follows from the fact that the prod-
uct in the category of sets is the Cartesian product.
Definition 106 If K is a context and if for every j ∈ J the map σj is an Sj -measure
of K, then the product measure
`
σ : K →aj∈J Sj
is defined by
σ(g) := (σj (g))j∈J . ♦
Hence, the preimages under σ are precisely such subsets that are intersections of
preimages under the measures σj . Each set of this kind is an extent of K, hence
σ is a scale measure.
We can slightly refine the argument in this proof: Every extent is the inter-
section of attribute extents. Hence, the preimages of extents of the product scale
under σ are precisely the intersections of preimages of the attribute extents of
Sj , j ∈ J. This leads to the following observation:
Proposition 134 The product measure is full if and only if the set of concepts of the form
is infimum-dense in B(K).
If all attributes of K are irreducible, the following holds true: The product
measure is full if and only if every attribute extent of K is the preimage of an
attribute extent under one of the scale measures σj , i.e., if for every attribute m
I
of K there exists some j ∈ J and some attribute mj ∈ Mj with m0 = σj−1 (mjj ).
Finally, we can use the notion of the product measure to give an alternative
definition of the derived context with respect to plain scaling:
Proposition 135 Let (G, M, W, I) be a complete many-valued context and let Sm , m ∈
M , be scales for the attributes of M . Furthermore, let K be the derived context with respect
to plain scaling. Then, for every many-valued attribute m ∈ M , the map
g 7→ m(g)
7.4 Measurability theorems 303
Proposition 135 has shown that full scale measures into semiproducts of scales
can be understood as a kind of inversion of plain scaling. Now, we can try to rec-
ognize derived contexts, i.e., to decide in the case of a given one-valued context
whether it could have been derived from a many-valued context through scaling
with given scales. Hence, the question is which contexts can be fully measured
in a semiproduct of nominal scales, ordinal scales etc. and, if so, which size the
necessary semiproduct must have. Proposition 133 is very useful in this context,
since it can be used to split up the problem. Therefore, we first examine how we
can recognize whether a given context allows a measure in one of the standard
scales. If this is the case, some of the attributes of the context can be combined to
form a many-valued attribute (with given scaling) and a repeated implementa-
tion of this procedure according to Proposition 133 finally yields a full measure
into the semiproduct.
Since every scale measure is surjective onto a subscale, it is useful to know the
subscales of the standard scales. In many cases they belong to the same family
of scales.
Proposition 136 The following families of scales have the property that every subscale
of a scale belonging to the family is equivalent to a scale of the same family:
f) cd
S = OP , if and only if there is a set P of extents with the following properties:
• The set P, ordered by set inclusion ⊆, is isomorphic to P.
• Every union of extents from P is an extent.
• For every object g ∈ G there is a largest extent Ug ∈ P which does not contain
g.
• P = {Ug | g ∈ G}.
g) S = CP , if and only if there is a set P of extents which satisfy the conditions under
f) and for which additionally the following statement is true:
• The complements of extents from P and the unions of such complements are
also extents.
Proof From Proposition 131 we know that K allows a surjective S-measure if
and only if there is a family US of extents of K and a map σ : G −→ GS such
that the (attribute) extents of S are precisely the images of US under σ. In other
words: In the system of the extents of K, those attribute extents of a scale must
occur which are isomorphic to S after clarification.
This makes a), b), c) and d) obvious. In order to show e), we first convince our-
selves of the fact that the complementary nominal scale has the property spec-
ified: For every scale value g ∈ GS , {g} is an extent, but also GS \ {g}. Hence,
the preimage sets of the scale values under a surjective S-measure form a par-
tition with the property specified in the Proposition. However, the converse is
also true: Given such a partition, a map of G onto GS mapping the classes of the
partition onto the values of the scale is an S-measure.
For f) we argue similarly: According to the definition, the extents of a contraor-
dinal scale are precisely the order ideals of P, the attribute extents are precisely
the complements of principal filters. Hence, the system of attribute extents of
S := Ocd P satisfies the conditions specified in f), and so do the preimages under
an S-measure. If, conversely, a system of such extents of a context is given, and
if ϕ is the order isomorphism of this system onto P, then we obtain a S-measure
through σ(g) := ϕ(Ug ) for all g ∈ G. Under these premises, the preimage of the
attribute extent {x ∈ P | x 6≥ p} is equal to
principal filters are convex sets. Hence this condition is necessary. It remains to
be shown that it is also sufficient.
Hence, let P be a system of extents of K which satisfies the conditions under
f), and let σ be the OP cd
-measure constructed in the proof of f). We shall prove
that under the additional condition the same map σ is also an OcP -measure (from
which the statement follows).
For this purpose we define a set system Q := {Vg | g ∈ G} through
[
Vg := {G \ U | U ∈ P, g ∈ U },
and show that Q satisfies the conditions specified under f), namely for the order
dual to P. The additional condition in g) guarantees that every Vg and all unions
of such sets are extents. According to the definition, we have furthermore
Vg ⊆ Vh ⇔ {U ∈ P | g ∈ U } ⊆ {U ∈ P | h ∈ U }
⇔ {U ∈ P | g 6∈ U } ⊇ {U ∈ P | h 6∈ U }
⇔ {U ∈ P | U ⊆ Ug } ⊇ {U ∈ P | U ⊆ Uh }
⇔ Ug ⊇ Uh .
Proposition 137 For every family S of scales, one of the following alternatives holds:
1. Every context is fully S-measurable.
2. Every fully S-measurable context is fully nominally measurable.
Proof First, we show that every context K is fully ordinally measurable, even
fully {O2 }-measurable. This follows immediately from Proposition 133: If we de-
fine for every attribute m of K an O2 -measure σm through
306 7 Context comparison and conceptual measurement
1 if g I m
σm (g) := ,
2 if (g, m) ∈
/I
then, because of σm −1
(1) = m0 , the product measure is full.
In this argument we have only made use of the fact that in O2 there are two
objects g, h with g 0 ⊂ h0 and g 0 6= h0 , i.e., that the context is not atomistic in the
sense of the definition on Page 45. Hence, the first alternative only does not occur
in the case that all scales in S are atomistic. Therefore, it only remains to be shown
that every atomistic scale is fully nominally measurable itself. This follows from
Proposition 138 below.
Concept lattices of atomistic contexts are atomistic, and every atomistic com-
plete lattice is isomorphic to the concept lattice of an atomistic context. The con-
text property “atomistic” means precisely that the extents of the object concepts
form a partition of the object set. It is inherited by semiproducts and by the
preimages under full scale measures: A context which can be fully measured
into a semiproduct of atomistic scales has to be atomistic itself. A reduced con-
text is atomistic if and only if from (g, m) ∈
/ I, g . m always follows.
Definition 108 An extent U of a context K = (G, M, I) is called n-valent if G \ U
is the disjoint union of n−1, but not of fewer extents. If K is atomistic, every extent
has a valence. In this case we define the valence of a set of extents as the supre-
mum of the valences involved. The attribute-valence VM (K) of an atomistic con-
text is the valence of the set of attribute extents, provided that K is reduced. In the
general case we say that an atomistic context K has attribute-valence VM (K) ≤ n
if and only if there is an infimum-dense set of concepts of K whose extents all
have a valence ≤ n. ♦
Proposition 138 A context is fully n-valued nominally measurable if and only if it has
attribute-valence ≤ n. Every atomistic context is fully nominally measurable.
Proof Every Nn -measure of K induces a partition of the object set into no more
than n extents, i.e., all those extents have a valence ≤ n. Hence, according to
Proposition 133 we find, for every full measure into a semiproduct of Nn -scales,
an infimum-dense set of concepts whose extents all have a valence ≤ n. If, on
the other hand, K has such a set of concepts, it is possible, by means of the same
proposition, to construct a product measure with the desired properties: An Nn -
measure can be assigned to every partition of the object set in at most ≤ n classes
by mapping the classes of the extent partition onto different objects of the nom-
inal scale.
An atomistic context (G, M, I) has an attribute-valence ≤ |G|, i.e., it is certainly
fully N|G| -measurable.
There has been little investigation as to which “measurability classes” there
are within the class of the atomistic scales. A first clue can be obtained if we also
define a valence for objects g ∈ G: g has valence n, if there are n − 1 objects
g1 , g2 , . . . , gn−1 (but no more) with the property that g, g1 , . . . , gn−1 generate the
7.4 Measurability theorems 307
same concept pairwise: {g, g1 }00 = {g, g2 }00 . . . = {g1 , g2 }00 . . . = {gn−2 , gn−1 }00
6= g 00 . VG (K) denotes the supremum of the valences of objects of K.
Proposition 139 If S consists of atomistic scales and if n is a natural number, the fol-
lowing statement is true: Nn is fully S-measurable if and only if S contains a scale S of
the object-valence VG (S) ≥ n.
Proof Every -preserving map of B(Nn ), n > 2 into a lattice which is not injec-
W
tive maps two atoms onto comparable elements. Therefore, by Proposition 132,
every non-trivial measure of Nn , n > 2 into an atomistic context is injective and
thus full. If N2 is measured fully into a semiproduct of atomistic scales, then at
least one of the factors must separate the two objects, and we have: If Nn is fully S-
measurable, then there is a scale S ∈ S, suchW that Nn is fully S-measurable. Again
by Proposition 132, then there must be a -embedding of B(Nn ) into B(S), in
the case of which object concepts are mapped onto object concepts, i.e., S has an
object-valence ≥ n.
Conversely, in a scale which has an object-valence ≥ n, we also find an object
of the valence n and thus a measure mapping Nn injectively and fully onto S.
From the last two propositions we draw a simple conclusion for a special case:
Proposition 140 If S consists of atomistic scales with an attribute-valence ≤ n and if S
contains a scale of the object-valence n (n ∈ N), then the following holds true: A context
K is fully S-measurable if it is fully n-valued nominally measurable.
This already suffices to provide us with an overview over the measurability
classes with respect to the atomistic standard scales. We have for all n ≥ 2
VG (Nn ) = VM (Nn ) = n,
VG (In ) = VM (In ) = 2,
VG (Ncn ) = VM (Ncn ) = 2.
From this follows:
Proposition 141 For a context K the following statements are equivalent:
1. K is fully dichotomially measurable.
2. K is fully interordinally measurable.
3. K is fully contranominally measurable.
The proposition also makes it possible to characterize the concept lattices of
many-valued contexts which are scaled plainly by means of elementary scales.
For this purpose, we need another definition: We say that an element x of an
atomistic complete lattice L has valence ≤ n if there is an n-element subset T
of L which contains x and which has the property that each atom of L is less
than or equal to precisely one element of T . If, in the special case n = 2, the set
T = {x, y} has the property specified, then we call y a pseudo-complement of
x.
308 7 Context comparison and conceptual measurement
7.1
Not only isomorphisms but also other classes of maps can be represented as con-
cepts, this has been worked out by W. Xia [425]. Theorem 60 has been taken from
[135], compare also Ganter and Reuter [155].
With regard to the group-theoretical background of Figure 7.1 extensive in-
formation can be found in Kerber ([217], in particular Chapter 3). M. Zickwolff
(from whose publications [428], [429] the example has been taken), has worked
out which information has to be entered into the diagram of the orbits of a (finite)
lattice (or, more generally, a relational structure) L with automorphism group
Γ , so that L can be reconstructed from it. Zickwolff’s approach focused on rule
exploration, about which more can be found in Section 6.4 of [154]. Later, the topic
was taken up, deepened and expanded by D. Borchmann [59]. The presentation
in this book is based on [62], where more detailed examples and further results
can be found.
Borchmann [59] also introduces context foldings and has GAP-programs for
computing concept lattice foldings directly from such context foldings. T. Schlem-
mer [334] suggests to further simplify the annotation of lattice foldings by using
group generators.
7.2
Definition 102 follows, with slight modifications, the article [119] by Erné, from
which we have also taken Theorem 62 and Proposition 126 and which contains
a lot of additional information concerning this subject. The question to which
7.5 Notes, references, and trends 309
7.3
Conceptual scales and conceptual measuring have first been discussed in [156].
Many results of this section can be found in different formulations in books such
as that of Blyth and Janowitz [52]. Otherwise, this section and the following one
make use of results taken from [159].
Hanika & Hirth [184] generalize Definition 105 by dropping the condition of
equal scale values. They introduce a finer-coarser relation on the set of all scale
measures (which start from a fixed formal context K) by calling a scale measure
σ1 finer than σ2 if every extent that is the preimage of an extent under σ2 is also
the preimage of an extent under σ1 . When this quasi-order is factorized by the
induced equivalence, the result is the complete lattices of the conceptual scalings of
K, which is isomorphic to the lattice of closure systems contained in the closure
system of all extents of K. This opens up the possibility of navigating between
possible scalings without having to commit to a conceptual scale beforehand.
7.4
The idea of concepts as the building blocks of human thought has been the object
of philosophical theories for centuries, as has logic as the doctrine of the forms of
thought. Formal Concept Analysis provides mathematical tools for the concep-
tual analysis of concrete facts and thereby supports human thinking and com-
munication. Of course, the rules of formal logic apply here as well, but additional
tasks arise. Therefore R. Wille and his co-workers have worked out a “Contextual
Logic” which provides a formal framework for Immanuel Kant’s “three essential
main functions of thinking”,
concepts – judgments – conclusions.
Here we discuss only the simplest parts of Contextual Logic in detail, begin-
ning with contextual attribute logic in Section 8.1. It mainly provides language
tools that allow attributes to be combined using logical connectives. A mathemat-
ically minded reader might object that this is merely reformulated Propositional
Logic, and in fact, we connect here to G. Boole, who gave the impulse in the mid-
dle of the 19th century. However, the language of contemporary mathematical
logic is noticeably shaped by its applications to mathematics itself. Boole spoke
of signs, classes, and of a universe of discourse, which in our understanding corre-
sponds to attributes, concept extents, and to the notion of a formal context. This does
not mean that the remarkable results of mathematical logic could be ignored. We
will use them whenever appropriate, but in doing so we will maintain a way of
speaking that is close to the logic of concepts. Logical Concept Analysis by Ferre
and Ridoux [128] offers an approach that is closer to the language of mathemat-
ical logic.
Concepts are not true or false, and formal concepts have no truth values. It
is therefore not obvious how a negation for (formal) concepts could be formu-
lated. Concept lattices are usually not complemented and therefore do not allow
a negation operation with the usual properties. Nevertheless, negation is dis-
cussed in the theory of concepts, and also the case where concepts are in oppo-
sition to each other. Wille [412, 413] has worked out how this can be expressed
for formal concepts. To map negation and opposition of concepts into Formal
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 311
B. Ganter, R Wille, Formal Concept Analysis, https://doi.org/10.1007/978-3-031-63422-2_9
312 8 Contextual concept logic
In addition to the attributes in the set M , one can consider compound attributes
formed by iteratively combining the attributes in M using the symbols ¬, , and
W
. The interpretation of when an object has a compound attribute follows the
V
standard definitions of propositional logic1 , with
¬m representing the negation of an attribute m ∈ M , with the extent being
V (¬m) := G \ m ,
0 0
(G, M ∪ C, I ∪ IC )
(G, M, I) | (G, C, IC )
and therefore has all concept extents of K and usually more, depending on the
choice of C. Context enrichment thus increases expressiveness, but also the num-
ber of formal concepts. It is usually not a good idea to add many or even all pos-
sible compound attributes because this can easily lead to a drastic increase in the
size of the concept lattice.
We must be careful when stating that two compound attributes “mean the
same”: we call such attributes
• extensionally equivalent over (G, M, I) if they have the same extent in this
formal context (G, M, I) and
• globally equivalent if they have the same extent in every context with at-
tribute set M .
Proposition 142 Two compound attributes of an attribute set M are globally equivalent
if and only if they are extensionally equivalent in the formal context
(P(M ), M, 3).
The extent of a sequent (A, S) consists of all objects g which have at least one of
the attributes in S or do not have at least one of the attributes in A. This is why
we sometimes find the notation
^ _
A→ S
314 8 Contextual concept logic
more intuitive, especially when examining the interaction of such sequents with
implications.
We will only consider disjoint sequents with A ∩ S = ∅, because every non-
disjoint sequent is globally equivalent to >. Sequents are naturally ordered by
and in fact does the compound attribute A imply the compound attribute S
V W
if and only if (A, S) is all-extensional.
The clause logic of a formal context (G, M, I) is the set of all sequents which
are all-extensional in (G, M, I). A clause set, i.e., a set C of disjoint sequents over
M , is called regular if it satisfies the condition
In words: (A, S) belongs to C iff every full sequent which contains (A, S) belongs
to C. An immediate consequence is that two regular clause sets containing the
same full sequents must be equal.
Theorem 67 A clause set is regular iff it is the clause logic of some formal context. Up
to clarification and isomorphism, this context is unique.
Proof It is easy to see that the clause logic of any formal context satisfies the
regularity condition. To construct a formal context for a given regular clause set
C, let {At | t ∈ T } denote the set of all first components of full sequents in C, and
let
G := P(M ) \ {At | t ∈ T }.
8.1 Contextual attribute logic 315
G is the extent of the conjunction of these full sequents in the test context. There-
fore these sequents are exactly the ones which are all-extensional in (G, M, 6 3),
and the clause logic of this formal context must be equal to C. For the unique-
ness, see below.
Attribute logic can, of course, not distinguish between objects having the same
object intent, and also does not encode information about the object names. But
up to object clarification, any two formal contexts having the same clause logic
must be isomorphic, as is easily proven. Thus the clause logic of (G, M, I) com-
pletely describes the system of object intents, in the same way as the implication
logic describes the system of concept intents.
Note that the intersection of regular clause sets is regular, and that the regular
clause sets thus form a closure system. The corresponding context is, up to clari-
fication, just the subposition of the individual ones. A full sequent (X, M \ X) is
in the regular closure of (A, S) iff (A, S) ≤ (X, M \ X).
The free extent of the union C1 ∪ C2 of two clause sets C1 and C2 is just the in-
tersection of the free extents of the two. This allows to combine the results of
the proposition to give further characterizations, e.g., that the convex subsets of
P(M ) are precisely the free extents of inconsistencies and disjunctions, etc.
Cumulated clauses
How do you find the most compact clause set possible for a formal context?
This question is well studied in mathematical logic (topic: “prime implicants”).
However, one should not expect such an elegant solution as for the implications.
When calculating with clauses, one quickly encounters intractable problems, i.e.
problems wich are NP-complete. On the other hand, computer science has de-
veloped algorithms –precisely for such calculations– (e.g., “SAT-solver”) that are
breathtakingly fast in practice. For formal contexts of moderate size it is therefore
quite possible to work with their clause logic.
Take a look at Figure 1.4 (page 30), the concept lattice to “Living beings and
water”. The largest element is attribute concept to the attribute “needs water to
live”, but it is not an object concept. So the intent of this concept is not an object
intent, and it does not occur as a “context row”. Instead, we see that every object
has one of the attributes a: needs water to live, g: can move around, b: lives on land, c:
lives in water, or d: needs chlorophyll. Using the abbreviations we can translate that
to saying that the sequent
({a}, {b, c, d, g}),
which can also be written as
a → b ∨ c ∨ d ∨ g,
is all-extensional in the formal context. Repeating the same for the attribute con-
cept for the attribute b: lives on land, we find that every such living being (of the
formal context) lives in water, needs chlorophyll, or can move around and has limbs.
What we get is the rule
c → b ∨ d ∨ (g ∧ h),
which is not a clause, but a combination of two clauses.
Such cumulated clauses, as we call expressions of the form
^ _^
A→ At ,
t∈T
show some surprising similarities to implications and are sometimes more intu-
itive than clauses alone. Since they include clauses, they have the same expres-
sivity, which at the same time means that they are algorithmically more complex
to handle than implications. We refer to Chapter 5 of [154] for details.
8.2 Pre-, proto-, and semiconcepts 317
A⊆G B⊆M
preconcept A ⊆ B 0 ( ⇔ B ⊆ A0 ) “rectangle”
protoconcept A0 = B 00 ( ⇔ B 0 = A00 ) A ⊆ B 0 B ⊆ A0
t-semiconcept A = B0
preconcept
u-semiconcept B = A0
concept A = B 0 and B = A0 A0 = B 00 B 0 = A00
protoconcept
A = B0 B = A0
t-semiconcept u-semiconcept
formal concept
While preconcepts are very general structural elements (just some objects
together with some attributes incident to them), protoconcepts (and thus also
semiconcepts) are quite natural for human understanding. It is not unusual to
describe a concept by some of its attributes and objects, if that is enough to de-
fine it unambiguously. It is not necessary to know the entire concept extent and
concept intents for this purpose.
318 8 Contextual concept logic
The set of all preconcepts of a formal context K is denoted by V(K),2 the set
of all protoconcepts by P(K). There are two natural ways to order preconcepts.
One ordering relation, noted as ⊆2 , is simply the containment of rectangles:
The maximal elements with respect to this order are exactly the formal concepts,
and the protoconcepts are precisely those less or equal to exactly one maximal
element.
The second definition generalizes the subconcept–superconcept order to pre-
concepts:
(A1 , B1 ) v (A2 , B2 ) :⇔ A1 ⊆ A2 and B2 ⊆ B1 .
It is surprisingly easy to prove that the set of all preconcepts, when ordered by
v, is also a complete lattice, which is even distributive. The following proposi-
tion also shows that the preconcepts can efficiently be generated using the Next
closure algorithm.
Proposition 145 (from [417]) For a formal context K := (G, M, I), the ordered set
(V(K), v) of all preconcepts is a completely distributive complete lattice, which is iso-
morphic to the concept lattice of the formal context
and
2 The letter V (Fraktur V ) resembles the German syllable „vor“, which translates to “pre”.
8.2 Pre-, proto-, and semiconcepts 319
[ \
sup(At , Bt ) = ( At , Bt ).
t∈T
t∈T t∈T
Hence (V(K), v) is a complete sublattice of the direct product of the two pow-
erset lattices (P(G), ⊆) and (P(M ), ⊇) and therefore completely distributive.
This proves the first assertion of the proposition. For the second one we consider
the assignment
ι
(A, B) 7→ (A ∪ (M \ B), (G \ A) ∪ B).
It can easily be checked that (A∪(M \B), (G\A)∪B) is a formal concept of V(K)
whenever (A, B) is a preconcept of K, and that the mapping ι is one-to-one. To
see that it is also onto, consider an arbitrary formal concept (C, D) of V(K), and
let A := C ∩ G and B := D ∩ M . From the definition of the incidence relation of
V(K) one concludes that A × B ⊆ I, and thus that (A, B) is a preconcept of K.
Moreover, C = A ∪ (M \ D) and D = B ∪ (G \ C), but since M \ D = M \ B and
ι
G \ C = G \ A, this shows that (A, B) 7→ (C, D). It remains to show that ι is not
only a bijection, but also an order-isomorphism, and this is easy:
Example For the contranominal scale Ncn , which we abbreviate with the symbol
6= , the context for the preconcepts takes a striking special form,
6= 6=
.
× 6=
320 8 Contextual concept logic
old young
male female
father daughter
son mother
Figure 8.3 Preconcept lattice for the formal context of family members in Figure 8.2. The
black nodes represent the ten formal concepts. The labelling is not the standard one de-
fined in Section 1.1. Not all irreducible elements are labelled. But it allows, with the usual
reading rule, to read off the sets A and B for each preconcept (A, B). Most of the nodes
in this diagram correspond to preconcepts (A, B) for which A or B are empty (31 of 47).
These form the two 16-element Boolean sublattices, one containing the smallest and the
other containing the largest element. Eight of the remaining elements are formal concepts.
The other eight correspond to preconcepts of size 1 × 1. The semicircles mark the proper
semiconcepts. There are no other protoconcepts in this example.
We already know (see the text following Definition 32 on p. 44) that its concept
lattice is isomorphic to the order relation of the contranominal scale, and more-
over we can infer from Proposition 18 on p. 45, that this context is isomorphic to
the context sum of n copies of the standard context of the three element chain.
Indeed, the order relation of the powerset of an n-element set consists of 3n pairs
and is isomorphic to the n-th power of a three-element chain. >
8.2 Pre-, proto-, and semiconcepts 321
Proposition 145 does not precisely charaterize the preconcept lattices, but this
is not an open problem, because Burgmann and Wille give a Basic Theorem on
preconcept lattices. We omit the technical details and instead quote their charac-
terization without proof.
Theorem 68 (from [68]) The preconcept lattices are (up to isomorphism) the com-
pletely distributive complete lattices in which the supremum of all atoms is equal or
greater than the infimum of all coatoms.
The lattice of preconcepts is useful because it provides an approach in the the-
ory of Formal Concept Analysis. Well-known algorithms such as Next closure
can thus be applied to preconcepts. However, this lattice does not represent a
generalization of the concept lattice to preconcepts in that the arithmetic opera-
tions do not coincide. The concept lattice is indeed order-embedded in the lattice
of preconcepts, but not as a sublattice. In the preconcept lattice, join and meet of
formal concepts are not necessarily formal concepts again. Therefore, another al-
gebraic structure is introduced next, whose elements are again the preconcepts,
but with other algebraic operations.
Definition 112 The preconcept algebra V(K) of a formal context K is the set
¬
V(K) endowed with operations u, t, ¬, , ⊥, and >, defined as follows:
♦
Two things stand out immediately: First, each result of these operations is a semi-
concept, so the semiconcepts form a subalgebra of each preconcept algebra. And
second, for formal concepts meet and join coincide with the concept lattice op-
erations. Generally for preconcepts, however, these operations are not lattice op-
erations and, except in special cases, do not have the usual properties. Meet and
join are not idempotent, double negation is not identity, etc. For details we refer
to the literature, in particular to Wille [417], where the equational theory of such
algebras is studied and finally a complete characterization is proved in terms of
a Basic Theorem on preconcept algebras (and for protoconcept algebras as well).
322 8 Contextual concept logic
of type (2, 2, 1, 1, 0, 0) that satisfies all equations which hold in all protoconcept
algebras.3 A double Boolean algebra is called pure, if
x u x = x or x t x = x
(Du ; u, t, ¬, ⊥, >)
where ∧ and ∨ denote meet and join in the concept lattice, and 0 is the least and
1 the largest concept of K. ♦
Note that weak negation and weak opposition can be expressed in terms of
the operations in Definition 112: For each formal concept x := (A, B) one has
¬ ¬
xM = (¬x) t (¬x) and xO = ( x) u ( x).
Concept algebras are concept lattices with two additional complement opera-
tions. Therefore we speak of dicomplemented lattices. However, an algebraic defi-
nition is usually done by equations or at least quasi-equations, but these are not
yet clear. Therefore, for the time being, we define as follows:
324 8 Contextual concept logic
Can the class of dicomplemented lattices be defined by equations, and if so, what is
a defining set of equations?
Some simple laws of concept algebras are compiled in the next proposition,
which is easily derived from the definitions of weak negation and opposition:
Proposition 146 The following hold in every concept algebra for all elements x and y:
All six conditions can be formulated as equations (for example, (2) is equiva-
lent to xM ∨(x∨y)M = xM ). These six conditions therefore might give an equational
base for the theory of dicomplemented lattices, but it is unknown if they do. As
a precaution, therefore, lattices that meet these conditions are given a different
name.
Definition 116 A weakly dicomplemented lattice is a bounded lattice equipped
with two unary operations M and O, satisfying the conditions (1)–(3) and (1’)–
(3’), listed in Proposition 146. ♦
4 This definition is slightly different from and supersedes the one in [412].
5 but does not suffice, as was shown by Kwuida & Machida [255].
8.3 Concept algebras 325
Proposition 147 Each weakly dicomplemented lattice satisfies the following equations:
Partial results
At the time of the second edition of this book, the representation issues for
dicomplemented lattices are open. Remarkable advances by Kwuida, however,
provide clues.
The question whether the quasi-equational class of dicomplemented lattices
is definable by equations alone refers to an important result of universal algebra,
namely to a theorem of Birkhoff that a class of algebras (of fixed signature) is
equationally definable if and only if it is closed under forming subalgebras, direct
products and homomorphic images. We had already mentioned that concept
326 8 Contextual concept logic
algebras cannot form an equational class because the lattice property of being
complete cannot be expressed by equations. But at least parts of the conditions
are fulfilled:
Theorem 71 (Kwuida [252])
• Direct products of concept algebras are isomorphic to concept algebras.
• Complete subalgebras of concept algebras are concept algebras.
• Homomorphic images of doubly founded concept algebras are isomorphic to concept
algebras.
For a proof we refer to the cited literature.
The weakly dicomplemented lattices form an equational class. We do not
know whether this is identical to the class of dicomplemented lattices. If this
can be shown, then the first open problem is also answered.
Note that a given lattice may have many weak dicomplementations. Kwuida
studied this for the case of the direct product of a two-element chain with an
n-element chain and showed that this lattice carries n2 weak dicomplementa-
tions, all of which are indeed dicomplementations. In other words, for each of
the n2 weak dicomplementations of this 2n-element lattice L, there exists a for-
mal context K whose concept algebra A(K) is isomorphic to L, equipped with
this dicomplementation.
Kwuida shows that this extends to arbitrary finite distributive lattices:
Theorem 72 (Kwuida [252]) Every finite distributive weakly dicomplemented lattice
is isomorphic to a concept algebra.
The example chosen to introduce the topic of this section comes (with some mod-
ifications) from the dissertation by S. Rudolph [323]. There it was used to sug-
gest the possibility of process exploration, i.e., of applying attribute exploration to
processes.
Example (Media player) A simple media player is considered. As visible at-
tributes, symbols are used to indicate which of six states the player is in:
Figure 8.4 shows a formal context with some obvious attributes for these states,
along with its simple concept lattice.
However, it is possible to influence the states. The media player provides four
commands for this purpose, which are also represented by symbols:
8.4 Contextual logic with relations 327
not playing
forwarding
rewinding
playing
idle
not playing
× ×
idle rewinding forwarding playing
×
× ×
× ×
× ×
× ×
which is not an extent of the formal context in Figure 8.4. We have thus intro-
duced a new attribute that can be called “ playing after ”, and can be symbol-
ized by
.playing
and which is defined to have the extension
0
.playing = {g ∈ G | changes g to h and (h, playing) ∈ I }.
This is, however, too special and only works, because is functional, meaning
that h is uniquely determined by g. The command is different, since the last
328 8 Contextual concept logic
Command given
or
or
or
Figure 8.5 The effect of the possible commands on the state of the media player.
column in Figure 8.5 in some cases has more than one entry per row. We then
must be more precise and specify, which of the possible outcomes is meant. For
a simpler notation, let us write
g h
to express that g, h ∈ G are objects such that h is one of the possible outcomes
when the command is given while the player is in state g. We introduce two
possibilities, ∃ .m and ∀ .m, with the extensions
0
∃ .m := {g ∈ G | ∃h∈G (g h and (h, m) ∈ I)}
0
∀ .m := {g ∈ G | ∀h∈G (g h implies (h, m) ∈ I)}.
0
In the above example, ∃ .idle consists of all states, as the media player al-
ways enters an idle state after a sufficiently long wait. After a short wait, however,
it may be that the previous state of playing, rewinding or forwarding has not yet
been completed. Thus
0
∀ .idle = , ,
in this example. Adding such attributes for all possible combinations can enlarge
the concept lattice considerably, even in this tiny example, see Figure 8.6. >
8.4 Contextual logic with relations 329
not playing
∀ .rewinding
∀ .f orwarding
∀ .idle ∀ .playing
idle
∀ .idle
rewinding playing
Figure 8.6 The concept lattice of the formal context from Figure 8.4, extended by relational
compound attributes. Only one name is given to each irreducible attribute.
In the example, we talked about “states” and “commands” that lead from one
state to another. However, the approach of relational concept analysis is more
general and not only suitable for processes, but only requires that additional re-
lations are given on the objects of a formal context. Recall the very first definition
in this book (Definition 1 on p. 1). There we had explained what a binary relation
on a set G is, namely a subset of G × G, i.e., a set of pairs of elements of G. This
can be generalized to k-ary relations on G, which are subsets R ⊆ Gk , i.e., sets
of k-tuples with entries from G.
Definition 117 A formal context K := (G, M, I) together with6 a set R of rela-
tions on G is a relational formal context. ♦
Contextual logic handles relations of all arities, but binary relations, i.e., relations
of arity 2, are the most frequently used ones. In this section we therefore only
deal with the case of binary relations and call them roles, which follows the way
of speaking that is common in Description Logics. Thus, a role is the same as a
binary relation on the object set.
For a relational context, further compound attributes can be defined in addi-
tion to those discussed in Section 8.1:
6 Admittedly, the definition remains somewhat vague at this point, since we do not specify a
notation for relational contexts. However, this is intentional, as there are different notations in
the literature, see Section 8.5.
330 8 Contextual concept logic
∀has-topping.vegetarian.
are allowed (with the parenthesis usually omitted). Staying with the pizza ex-
ample: If one of the toppings contains bacon, i.e., if the pizza has the attribute
∃has-topping.∃contains.bacon
“test context” for terminological attributes, and thus no obvious way to decide
whether two such attributes are globally equivalent, an indispensable prerequi-
site for attribute exploration.
At this point, however, Formal Concept Analysis can benefit from the results
obtained for description logics. There, the satisfiability problem for the respective
logic is studied both theoretically and algorithmically. Many decision questions
that arise for terminological attributes can be reformulated as satisfiability prob-
lems, e.g. the question whether such an attribute follows (globally) from other
attributes. Typical results characterize the computational complexity of satisfia-
bility and the development of powerful reasoners, which can solve the problem
effectively. The cited authors have therefore pursued the goal of establishing con-
nections from Formal Concept Analysis to Description Logics that make these
reasoners accessible.
Sebastian Rudolph has pointed out (in personal communication) that it is per-
fectly possible to construct relational test contexts using description logics. How-
ever, these are infinite and inferior to reasoners for testing global equivalence.
8.1
8.2
The notion of preconcepts occurs in Stahl & Wille [348], that of semiconcepts in
Luksch & Wille [268], and protoconcepts are discussed in [412]. Proofs of the
“basic theorems” for semi- and for protoconcept lattices can be found in Vorm-
brock & Wille [381], for preconcept lattices in Burgmann & Wille [68]. Figure 8.3
is from [417].
8.3
Concept algebras were introduced in [412], where also a series of results can be
found. Later results are largely contained in Kwuida’s dissertation [252] and the
publications based on it.
8.5 Notes, references, and trends 333
8.4
Research developments
In 1996, the first version of this book (in German) made it visible that by then an
extensive mathematical basis of Formal Concept Analysis had been worked out.
Rudolf Wille therefore turned more to the question of how this theory could
be embedded in a larger framework and began to develop a contextual logic.
Here, “logic” is to be understood as the doctrine of the forms of human thought,
and the addition of “contextual” refers to the goal of uncovering the regulari-
ties of the data at hand in each case, of the “context”. Wille was referring to the
philosopher Immanuel Kant, who taught elementary logic as “the theory of the
three main essential functions of thinking – concepts, judgments, and conclusions”,
with concepts as the basic units of thought, judgments as combinations of con-
cepts and conclusions as entailments between judgments, see [414] and Predi-
ger [305]. Was it possible to complement this philosophical approach, developed
over centuries, with a theory of formal concepts, formal judgments, and formal
conclusions? To mathematize the formal judgments, concept graphs were intro-
duced, a mathematically adapted variant of J. Sowa’s conceptual graphs [346].
A first systematic presentation of the approach is Prediger’s dissertation [304].
However, a comprehensive account of contextual logic does not yet exist, except
for the condensed “summary” in [414].
Description Logics (DL) are a family of logics that restrict the full expressive
power of predicate logic in favor of practicality for knowledge representation
and knowledge processing. There is far advanced theory and extremely effective
algorithms for this. They are also the basis of the web ontology language OWL.
And they offer a natural language for a relational extension of Formal Concept
Analysis. There was early interest in combining description logics and Formal
Concept Analysis. F. Baader, editor of the Description Logic Handbook [8], used
FCA methods as early as 1995 in his research [7]. Together with his students
and co-authors, he has consistently continued these efforts. This has resulted in
numerous publications with substantial results. Of particular note are the disser-
tations by S. Rudolph [323] B. Sertkaya [340], F. Distel [97], D. Borchmann [60],
and F. Kriegel [232, 233], but also Baader & Molitor [10], Baader & Sertkaya [11],
Baader et al. [9]. Other contributions come e.g. from Stumme [360].
Conversely, there are numerous efforts to enrich Formal Concept Analysis
with elements of DL languages. Prediger & Stumme [307] use DL for introduc-
ing logical scaling, and Prediger [306] works out a terminological attribute logic.
334 8 Contextual concept logic
Rouane-Hacene et al. [322, 196] suggest to combine methods from the two fields.
An early paper on formal contexts with a relational structure is due to Priss [308].
For a contextual logic, as introduced by R. Wille, the standard version of For-
mal Concept Analysis mathematizes the part of Contextual Concept Logic. But its
expressive power of the Formal Concept Analysis is not sufficient for Contex-
tual Judgment Logic of Conclusion Logic. To capture the logical content of concept
graphs, Wille [415] introduced power context families, which encode relations
of arbitrary arity.
Other authors chose different approaches. Ferre & al. [130] use predicate logic,
Ferre & Cellier introduce Graph-contexts for which the object set is replaced by
words of objects to describe knowledge graphs (both formalisms allow for re-
lations of arbitrary arity). Several authors restrict to binary relations. Rudolph
uses binary power context families, Huchard and her coauthors, in development
of a relational concept analysis (see [196] for an early and [290] for a more re-
cent contribution), use relational context families, where each relation is given by
its incidence matrix and has a domain and a range. Their data type is similar to
what Dörflein & Wille [104] call a coherence network and for which they have even
proved a “Basic Theorem”.
For our presentation in Section 8.4, the choice of the data type is of secondary
importance.
References
[1] Albano, A.: Polynomial growth of concept lattices, canonical bases and
generators: extremal set theory in Formal Concept Analysis. Ph.D. thesis,
TU Dresden (2017). Qucosa Dresden → 113, 250, 279
[2] Albano, A., Chornomaz, B.: Why concept lattices are large: extremal the-
ory for generators, concepts, and VC-dimension. International Journal of
General Systems 46(5), 440–457 (2017). → 63, 251, 252
[3] Almazaydeh, A., Behrisch, M., Vargas-García, E., Wachtel, A.: Arrow re-
lations in lattices of integer partitions. Preprint arXiv:2403.07217 (2024).
→ 71
[4] Almeida, J., Bordalo, G., Dwinger, P.: Lattices, Semigroups, and Universal
Algebra. Plenum Press, New York–London (1990). doi:10.1007/978-1-
4899-2608-1. → 348
[5] Armstrong, W.W.: Dependency structures of data base relationships. In:
IFIP congress, vol. 74, pp. 580–583. Geneva, Switzerland (1974). → 94
[6] Atif, J., Bloch, I., Distel, F., Hudelot, C.: Mathematical morphology oper-
ators over concept lattices. In [73] pp. 28–43 (2013). doi:10.1007/978-3-
642-38317-5_2. → 309
[7] Baader, F.: Computing a minimal representation of the subsumption lat-
tice of all conjunctions of concepts defined in a terminology. In: Proc. Intl.
KRUSE Symposium, August, pp. 168–178. Santa Cruz, USA (1995). → 333
[8] Baader, F., Calvanese, D., McGuinness, D.L., Nardi, D., Patel-Schneider,
P.F.: The Description Logic handbook: Theory, implementation and
applications. Cambridge University Press (2nd edition, 2007).
doi:10.1017/CBO9780511711787. → 333
[9] Baader, F., Ganter, B., Sattler, U., Sertkaya, B.: Completing description logic
knowledge bases using formal concept analysis. In: IJCAI, vol. 7, pp. 230–
235 (2007). doi:10.25368/2022.155. → 333
[10] Baader, F., Molitor, R.: Building and structuring description logic knowl-
edge bases using least common subsumers and concept analysis. In [152]
pp. 292–305 (2000). doi:10.1007/10722280_20. → 333
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 335
B. Ganter, R Wille, Formal Concept Analysis, https://doi.org/10.1007/978-3-031-63422-2
336 References
[11] Baader, F., Sertkaya, B.: Applying formal concept analysis to description
logics. In [114] pp. 261–286 (2004). doi:10.1007/978-3-540-24651-0_24.
→ 333
[12] Babin, M.A., Kuznetsov, S.O.: Recognizing pseudo-intents is coNP-
complete. In [242] pp. 294–301 (2010). → 115
[13] Baixeries, J.: Lattice characterization of Armstrong and symmetric depen-
dencies. Ph.D. thesis, Universitat Politècnica de Catalunya, Spain (2007).
→ 115
[14] Baixeries, J.: Database dependencies and formal concept analysis. Preprint
arXiv:2403.13914 (2024). → 115
[15] Baixeries, J., Codocedo, V., Kaytoue, M., Napoli, A.: Three views on de-
pendency covers from an FCA perspective. In [111] pp. 78–94 (2023).
doi:10.1007/978-3-031-35949-1_6. → 115
[16] Baker, K.A., Wille, R.: Lattice Theory and Its Applications: In Celebration
of Garrett Birkhoff’s 80th Birthday. Heldermann Verlag (1995). → 343,
357
[17] Balbes, R., Dwinger, P.: Distributive lattices. Journal of Symbolic Logic
42(4) (1977). doi:10.2307/2271885. → 253
[18] Balcázar, J.L., Baixeries, J.: Characterizations of multivalued dependencies
and related expressions. In: Discovery Science: 7th International Confer-
ence, DS 2004, Padova, Italy, October 2-5, 2004. Proceedings 7, pp. 306–313.
Springer (2004). doi:10.1007/978-3-540-30214-8_25. → 115
[19] Baldy, P., Mitas, J.: Generalized dimension of an ordered set and its Mac-
Neille completion. Order 11(2), 135–148 (1994). doi:10.1007/bf01108598.
→ 281
[20] Banaschewski, B.: Hüllensysteme und Erweiterung von Quasi-
Ordnungen. Mathematical Logic Quarterly 2(8-9), 117–130 (1956).
doi:10.1002/malq.19560020803. → 68, 281
[21] Banaschewski, B., Bruns, G.: Categorical characterization of the MacNeille
completion. Arch. Math. 18, 369–377 (1967). doi:10.1007/bf01898828.
→ 71
[22] Bandelt, H.J.: Tolerance relations on lattices. Bulletin of
the Australian Mathematical Society 23(3), 367–381 (1981).
doi:10.1017/s0004972700007255. → 160, 214
[23] Barbut, M.: Note sur l’algèbre des techniques d’analyse hiérarchique. B.
Matalon (1965). doi:10.1515/9783111540924-010. → 68
[24] Barbut, M., Monjardet, B.: Ordre et classification, vols. 1 and 2. Hachette,
Paris, France (1970). → 68
[25] Bartenschlager, G.: Free bounded distributive lattices generated by finite
ordered sets. Ph.D. thesis, Technische Hochschule Darmstadt (1994).
→ 71, 247
[26] Bartenschlager, G.: Free bounded distributive lattices over finite ordered
sets and their skeletons. Acta Math. Univ. Comenianae 64(1), 1–23 (1995).
→ 71
References 337
[27] Bartl, E., Krupka, M.: Residuated lattices of block relations: size reduction
of concept lattices. International Journal of General Systems 45(7-8), 773–
789 (2016). doi:10.1080/03081079.2016.1194021. → 160
[28] Barwise, J., Seligman, J.: Information flow: the logic of distributed sys-
tems. Cambridge University Press (1997). doi:10.1017/cbo9780511895968.
→ 75, 296
[29] Bazin, A.: A triadic generalisation of the Boolean concept lattice. In [111]
pp. 95–105 (2023). → 73
[30] Bazin, A., Beaudou, L., Kahn, G., Khoshkhah, K.: Bounding the num-
ber of minimal transversals in tripartite 3-uniform hypergraphs. Discrete
Mathematics & Theoretical Computer Science 23 (special issues) (2023).
doi:10.46298/dmtcs.7129. → 73
[31] Bazin, A., Kahn, G.: Reduction and introducers in d-contexts. In [82] pp.
73–88 (2019). doi:10.1007/978-3-031-35949-1_7. → 73
[32] Behrisch, M., Chavarri Villarello, A., Vargas-García, E.: Representing par-
tition lattices through FCA. In [66] (2021). doi:10.1007/978-3-030-77867-
5_1. → 71
[33] Belohlavek, R.: Fuzzy relational systems: foundations and principles,
vol. 20. Kluwer Academic/Plenum Publishers, New York (2002).
doi:10.1108/k.2003.06732iae.005. → 73
[34] Bĕlohlávek, R.: Concept lattices and order in fuzzy logic. An-
nals of pure and applied logic 128(1-3), 277–298 (2004).
doi:10.1016/j.apal.2003.01.001. → 73
[35] Belohlavek, R., Glodeanu, C., Vychodil, V.: Optimal factorization of three-
way binary data using triadic concepts. Order 30, 437–454 (2013).
doi:10.1007/s11083-012-9254-4. → 281
[36] Belohlavek, R., Trnecka, M.: Basic level of concepts in formal concept ana-
lysis. In [102] pp. 28–44 (2012). doi:10.1007/978-3-642-29892-9_9. → 279
[37] Belohlavek, R., Trnecka, M.: Basic level in formal concept analysis: Inter-
esting concepts and psychological ramifications. In: Twenty-Third Inter-
national Joint Conference on Artificial Intelligence (2013). → 279
[38] Belohlavek, R., Trnecka, M.: From-below approximations in Boolean ma-
trix factorization: Geometry and new algorithm. Journal of Computer and
System Sciences 81(8), 1678–1697 (2015). doi:10.1016/j.jcss.2015.06.002.
→ 281
[39] Belohlavek, R., Trnecka, M.: Basic level of concepts in formal concept ana-
lysis 1: formalization and utilization. International Journal of General Sys-
tems 49(7), 689–706 (2020). doi:10.1080/03081079.2020.1837794. → 279
[40] Belohlavek, R., Vychodil, V.: Discovery of optimal factors in binary data
via a novel method of matrix decomposition. Journal of Computer and
System Sciences 76(1), 3–20 (2010). doi:10.1016/j.jcss.2009.05.002. → 281
[41] Belohlavek, R., Vychodil, V.: Factorizing three-way binary data with triadic
formal concepts. In: International Conference on Knowledge-Based and
Intelligent Information and Engineering Systems, pp. 471–480. Springer
(2010). doi:10.1007/978-3-642-15387-7_51. → 73
338 References
[42] Belohlavek, R., Vychodil, V.: Formal concept analysis and linguistic
hedges. International Journal of General Systems 41(5), 503–532 (2012).
doi:10.1080/03081079.2012.685936. → 73
[43] Bennett, M.K., Birkhoff, G.: Two families of Newman lattices. Algebra Uni-
versalis 32, 115–144 (1994). doi:10.1007/bf01190819. → 235
[44] Benthem, J.: Information transfer across Chu spaces. Logic Journal of the
IGPL 8(6), 719–731 (2000). doi:10.1093/jigpal/8.6.719. → 309
[45] Berry, A., Sigayret, A.: Representing a concept lattice by a
graph. Discrete Applied Mathematics 144(1-2), 27–42 (2004).
doi:10.1016/j.dam.2004.02.016. → 71
[46] Berry, A., Sigayret, A.: Dismantlable lattices in the mirror. In [73] pp. 44–59
(2013). doi:10.1007/978-3-642-38317-5_3. → 70, 161
[47] Bertet, K., Borchmann, D., Cellier, P., Ferré, S. (eds.): Proceedings of the
14th ICFCA, vol. LNAI 10308. Springer (2017). doi:10.1007/978-3-319-
59271-8. → 344, 348
[48] Biedermann, K.: A foundation of the theory of trilattices. Ph.D. thesis, TU
Darmstadt (1998). → 73
[49] Birkhoff, G.: Lattice theory, vol. 25. American Mathematical Soc. (1940).
doi:10.1090/coll/025. → 20, 68, 251
[50] Birkhoff, G.: Lattice Theory 3rd. ed. American Mathematical Soc. (1967).
→ 20
[51] Birkhoff, G., Bennett, M.K.: The convexity lattice of a poset. Order 2, 223–
242 (1985). doi:10.1007/bf00333128. → 71
[52] Blyth, T.S., Janowitz, M.F.: Residuation theory. Pergamon Press (1972).
doi:10.1016/b978-0-08-016408-3.50004-8. → 20, 309
[53] Bock, H.H.: Classification and related methods of data analysis. Dis-
tributors for the USA and Canada, Elsevier Science Pub. Co. (1988).
doi:10.5771/0943-7444-1989-2-115. → 351, 358
[54] Bock, H.H., Ihm, P.: Classification, Data Analysis, and Knowledge Organi-
zation: Models and Methods with Applications. Springer Science & Busi-
ness Media (2012). doi:10.1007/978-3-642-76307-6. → 351, 356, 357
[55] Bock, H.H., Lenski, W., Richter, M.M.: Information Systems and Data Ana-
lysis: Prospects-Foundations-Applications. Springer Science & Business
Media (2013). doi:10.1007/978-3-642-46808-7. → 349, 357
[56] Bock, H.H., Polasek, W. (eds.): Data Analysis and Information Systems:
Statistical and Conceptual Approaches. Springer Science & Business Me-
dia, University of Basel (1996). doi:10.1007/978-3-642-80098-6. → 354,
356
[57] Bokowski, J., Kollewe, W.: On representing contexts in line arrangements.
Order 8, 393–403 (1991). doi:10.1007/bf00571189. → 113
[58] Boley, M., Grosskreutz, H., Gärtner, T.: Formal concept sampling for
counting and threshold–free local pattern mining. In: Proc. of
the SIAM Int. Conf. on Data Mining (SDM 2010). SIAM (2010).
doi:10.1137/1.9781611972801.16. → 69
References 339
[74] Cellier, P., Ducassé, M., Ferré, S., Ridoux, O.: Formal concept analysis
enhances fault localization in software. In [279] pp. 273–288 (2008).
doi:10.1007/978-3-540-78137-0_20. → 74
[75] Chaudron, L., Maille, N.: Generalized formal concept analysis. In [152]
pp. 357–370 (2000). doi:10.1007/10722280_25. → 75
[76] Chornomaz, B.: Algebraicity and the tensor product of concept lattices.
In [173] pp. 54–66 (2014). doi:10.1007/978-3-319-07248-7_5. → 214
[77] Codd, E.F.: A relational model of data for large shared data
banks. Communications of the ACM 13(6), 377–387 (1970).
doi:10.7551/mitpress/12274.003.0034. → 71
[78] Cogis, O.: Le dimension Ferrers des graphs orientés. Ph.D. thesis, Univer-
sité Curie, Paris (1980). → 281
[79] Cogis, O.: On the Ferrers dimension of a digraph. Discrete Mathematics
38(1), 47–52 (1982). doi:10.1016/0012-365x(82)90167-4. → 281
[80] Colomb, P., Irlande, A., Raynaud, O.: Counting of Moore families for n=
7. In [256] pp. 72–87 (2010). doi:10.1007/978-3-642-11928-6_6. → 13, 20
[81] Crawley, P., Dilworth, R.P.: Algebraic theory of lattices. Prentice Hall
(1973). → 278, 280
[82] Cristea, D., Le Ber, F., Sertkaya, B. (eds.): Proceedings of the 15th ICFCA.
Springer (2019). doi:10.1007/978-3-030-21462-3. → 337, 344, 349, 352
[83] Czédli, G.: Factor lattices by tolerances. Acta Sci. Math.(Szeged) 44(1-2),
35–42 (1982). → 160
[84] Davey, B.A., Priestley, H.A.: Introduction to lattices and order. Cambridge
University Press (2002). doi:10.1017/cbo9780511809088. → 20
[85] Day, A.: A simple solution to the word problem for lattices. Canadian
Mathematical Bulletin 13(2), 253–254 (1970). doi:10.4153/cmb-1970-051-
0. → 247
[86] Day, A.: Characterizations of finite lattices that are bounded-homomor-
phic images or sublattices of free lattices. Canadian Journal of Mathemat-
ics 31(1), 69–78 (1979). doi:10.4153/cjm-1979-008-x. → 70, 247, 280
[87] Day, A., Herrmann, C.: Gluings of modular lattices. Order 5, 85–101
(1988). doi:10.1007/bf00143900. → 214
[88] Day, A., Nation, J., Tschantz, S.: Doubling convex sets in lattices and
a generalized semidistributivity condition. Order 6, 175–180 (1989).
doi:10.1007/bf02034334. → 247, 280
[89] Degens, P.O., Hermes, H.J., Opitz, O. (eds.): Die Klassifikation und ihr
Umfeld. Indeks–Verlag, Frankfurt (1986). doi:10.5771/0943-7444-1987-2-
104. → 344
[90] Deiters, K., Erné, M.: Negations and contrapositions of complete lat-
tices. Discrete Mathematics 181(1-3), 91–111 (1998). doi:10.1016/s0012-
365x(97)00047-2. → 70
[91] Deiters, K., Erné, M.: Sums, products and negations of contexts
and complete lattices. Algebra Universalis 60(4), 469–496 (2009).
doi:10.1007/s00012-009-2141-1. → 70
References 341
[92] Delugach, H.S., Stumme, G. (eds.): Proceedings of the 9th ICCS, vol.
LNAI 2120. Springer Science & Business Media (2001). doi:10.1007/3-540-
44583-8. → 344, 359
[93] Diday, E.: Introduction à l’approche symbolique en analyse des
données. RAIRO-Operations Research 23(2), 193–236 (1989).
doi:10.1051/ro/1989230201931. → 75
[94] Dilworth, R.: Proof of a conjecture on finite modular lattices. The Dilworth
Theorems: Selected Papers of Robert P. Dilworth pp. 219–224 (1990).
doi:10.1007/978-1-4899-3558-8_21. → 259
[95] Dilworth, R.P.: Lattices with unique irreducible decompositions. In: The
Dilworth Theorems: Selected Papers of Robert P. Dilworth, pp. 93–99.
Springer (1990). doi:10.1007/978-1-4899-3558-8_10. → 280
[96] Distel, F.: Hardness of enumerating pseudo-intents in the lectic order.
In [256] pp. 124–137 (2010). doi:10.1007/978-3-642-11928-6_9. → 115
[97] Distel, F.: Learning description logic knowledge bases from data using
methods from formal concept analysis. Ph.D. thesis, TU Dresden (2011).
Qucosa Dresden, → 331, 333
[98] Distel, F., Sertkaya, B.: On the complexity of enumerating
pseudo-intents. Discrete Applied Mathematics 159(6) (2011).
doi:10.1016/j.dam.2010.12.004. → 115
[99] Doerfel, S.: A context-based description of the doubly founded concept
lattices in the variety generated by M3 . In [372] pp. 93–106 (2011).
doi:10.1007/978-3-642-20514-9_9. → 70, 259, 280
[100] Doignon, J.P., Ducamp, A., Falmagne, J.C.: On realizable biorders and the
biorder dimension of a relation. Journal of Mathematical Psychology
28(1), 73–109 (1984). doi:10.1016/0022-2496(84)90020-8. → 281
[101] Doignon, J.P., Falmagne, J.C.: Knowledge spaces. Springer Science & Busi-
ness Media (1999). doi:10.1007/978-3-642-58625-5. → 75, 315
[102] Domenach, F., Ignatov, D., Poelmans, J. (eds.): Proceedings of the 10th
ICFCA, vol. LNAI 7278. Springer (2012). doi:10.1007/978-3-642-29892-9.
→ 337, 344, 349, 354
[103] Domenach, F., Leclerc, B.: Biclosed binary relations and Galois connec-
tions. Order 18, 89–104 (2001). doi:10.1023/a:1010662327346. → 161
[104] Dörflein, S.K., Wille, R.: Coherence networks of concept lattices: the basic
theorem. In [149] pp. 344–359 (2005). doi:10.1007/978-3-540-32262-7_24.
→ 334
[105] Dorn, G., Frank, R.D., Ganter, B., Kipke, U., Poguntke, W., Wille, R.:
Forschung und Mathematisierung — Suche nach Wegen aus dem Elfen-
beinturm. Wechselwirkung 15, 20–23 (1982). → 68
[106] Dorninger, D., Eigenthaler, G., Kaiser, H.K., Müller, W.B. (eds.): Contri-
butions to general algebra, vol. 7. Hölder–Pichler–Tempsky, Wien (1991).
→ 355, 359, 360
[107] Dubois, D., Prade, H.: Possibility theory and formal concept analysis in
information systems. In: International Fuzzy Systems Association World
Congress and Conference of the European Society for Fuzzy Logic and
342 References
[145] Ganter, B.: “Properties of Finite Lattices” by S. Reeg and W. Weiß, re-
visited. In memoriam Peter Burmeister (1941–2019). In [82] pp. 99–109
(2019). doi:10.1007/978-3-030-21462-3_8. → 98, 353
[146] Ganter, B.: Notes on integer partitions. International Journal of Approxi-
mate Reasoning 142, 31–40 (2022). doi:10.1016/j.ijar.2021.11.004. → 71
[147] Ganter, B., Glodeanu, C.V.: Ordinal factor analysis. In [102] pp. 128–139
(2012). doi:10.1007/978-3-642-29892-9_15. → 281
[148] Ganter, B., Glodeanu, C.V.: Factors and skills. In [173] pp. 173–187 (2014).
doi:10.1007/978-3-319-07248-7_13. → 281
[149] Ganter, B., Godin, R. (eds.): Proceedings of the 3rd ICFCA 2005, vol.
LNAI 3403. Springer (2005). doi:10.1007/b105806. → 341, 355, 357
[150] Ganter, B., Krauße, R.: Pseudo-models and propositional Horn in-
ference. Discrete Applied Mathematics 147(1), 43–55 (2005).
doi:10.1016/j.dam.2004.06.019. → 107, 114
[151] Ganter, B., Kuznetsov, S.O.: Pattern structures and their projections.
In [92] pp. 129–142 (2001). doi:10.1007/3-540-44583-8_10. → 74
[152] Ganter, B., Mineau, G. (eds.): Proceedings of the 8th ICCS 2000, vol.
LNAI 1867. Springer (2000). doi:10.1007/10722280. → 335, 340, 343,
359
[153] Ganter, B., Nevermann, P., Reuter, K., Stahl, J.: How small can a lattice of
order-dimension n be? Order 3, 345–353 (1987). doi:10.1007/bf00340776.
→ 281
[154] Ganter, B., Obiedkov, S.: Conceptual exploration. Springer (2016).
doi:10.1007/978-3-662-49291-8. → 74, 77, 82, 94, 98, 99, 106, 112, 113, 114,
115, 308, 316, 331
[155] Ganter, B., Reuter, K.: Finding all closed sets: A general approach. Order
8, 283–290 (1991). doi:10.1007/bf00383449. → 112, 308
[156] Ganter, B., Stahl, J., Wille, R.: Conceptual measurement and many valued
contexts. In [166] (1986). → 309
[157] Ganter, B., Stumme, G., Wille, R.: Formal concept analysis: foundations
and applications, vol. 3626. Springer (2005). doi:10.1007/978-3-540-31881-
1. → 339, 349, 357
[158] Ganter, B., Wille, R.: Implikationen und Abhängigkeiten zwischen Merk-
malen. In [89] pp. 171–185 (1986). English version in [47]. → 113, 114
[159] Ganter, B., Wille, R.: Conceptual scaling. In: F.S. Roberts (ed.) Appli-
cations of combinatorics and graph theory to the biological and social
sciences, pp. 139–167. Springer (1989). doi:10.1007/978-1-4684-6381-1_6.
→ 71, 114, 309
[160] Ganter, B., Wille, R.: Subtensorial decompositions of concept lattices.
Tech. rep., Bericht MATH-AL-1-1994, TU Dresden FB Mathematik (1994).
→ 214
[161] Ganter, B., Wille, R.: Formale Begriffsanalyse – Mathematische Grundla-
gen. Springer-Verlag (1996). doi:10.1007/978-3-642-61450-7. → vii, 345
References 345
[162] Ganter, B., Wille, R.: Contextual attribute logic. In: International
Conference on Conceptual Structures, pp. 377–388. Springer (1999).
doi:10.1007/3-540-48659-3_23. → 332
[163] Ganter, B., Wille, R.: Formal Concept Analysis – Mathematical Founda-
tions. Springer-Verlag (1999). doi:10.1007/978-3-642-59830-2. Translated
from [161]. → vii, 345
[164] Ganter, B., Wille, R.: Xingshi Gainian Fenxi. Science Press (2005). Trans-
lated from [163]. → vii
[165] Ganter, B., Wille, R., Wolff, K.E. (eds.): Beiträge zur Begriffsanalyse. B.I.-
Wissenschaftsverlag Mannheim (1987). → 342, 343, 350, 358
[166] Gaul, W., Schader, M. (eds.): Proceedings of the 9th Annual Meeting of
the Classification Society, Classification as a Tool of Research, vol. 9. North
Holland, University of Karlsruhe, Germany (1986). → 344, 351, 355
[167] Gepperth, U.: Automatisches Zeichnen begriffsanalytischer Liniendi-
agramme unter Verwendung flußalgorithmischer Berechnungen von
Merkmalsketten. Diplomarbeit, FH Darmstadt (1990). → 113
[168] Geyer, W.: Generalizing semidistributivity. Order 10, 77–92 (1993).
doi:10.1007/bf01108710. → 280
[169] Geyer, W.: On context patterns associated with concept lattices. Order
10(4), 363–373 (1993). doi:10.1007/bf01108830. → 70
[170] Geyer, W.: The generalized doubling construction and formal concept ana-
lysis. Algebra Universalis 32, 341–367 (1994). doi:10.1007/bf01235175.
→ 247
[171] Geyer, W.: On Tamari lattices. Discrete Mathematics 133(1-3), 99–122
(1994). doi:10.1016/0012-365x(94)90019-1. → 235, 247
[172] Glodeanu, C.V.: Triadic factor analysis. In [242] pp. 127–138 (2010). → 73,
281
[173] Glodeanu, C.V., Kaytoue, M., Sacarea, C. (eds.): Proceedings of the 12th
ICFCA, vol. LNAI 8478. Springer (2014). doi:10.1007/978-3-319-07248-7.
→ 340, 344, 354
[174] Grätzer, G.: On the complete congruence lattice of a complete lattice with
an application to universal algebra. CR Math. Rep. Acad. Sci. Canada 11,
105–108 (1989). → 160
[175] Grätzer, G.: General lattice theory. Springer Science & Business Media
(2002). doi:10.1007/978-3-0348-9326-8. → 20, 70
[176] Greechie, R.: On generating pathological orthomodular structures. In:
Technical Report no. 13. Kansas State University (1970). → 247
[177] Greene, C., Markowsky, G.: A combinatorial test for local distributivity. In:
Notices of the American Mathematical Society, vol. 22 (2), pp. A299–A299.
AMS (1975). → 280
[178] Grötzsch, A.: Begriffsanalytische Darstellung verallgemeinerter Hüllen-
systeme. Diplomarbeit, TU Dresden (2005). → 71, 280, 291
[179] Guigues, J.L., Duquenne, V.: Familles minimales d’implications informa-
tives résultant d’un tableau de données binaires. Mathématiques et Sci-
ences humaines 95, 5–18 (1986). → 96, 97, 113
346 References
[214] Kent, R.E.: Rough concept analysis. In: Rough Sets, Fuzzy Sets and Knowl-
edge Discovery: Proceedings of the International Workshop on Rough Sets
and Knowledge Discovery (RSKD’93), Banff, Alberta, Canada, 12–15 Oc-
tober 1993, pp. 248–255. Springer (1994). doi:10.1007/978-1-4471-3238-
7_30. → 74
[215] Keprt, A.: Using blind search and formal concepts for binary factor analy-
sis. In: DATESO, vol. 98, pp. 128–140 (2004). → 281
[216] Keprt, A., Snášel, V.: Binary factor analysis with help of formal concepts.
In: CLA’04, vol. CEUR-WS 110, pp. 90–101 (2004). → 281
[217] Kerber, A.: Algebraic combinatorics via finite group actions. BI-
Wissenschaftsverlag, Mannheim (1991). → 308
[218] Kerber, A., Lex, W.: Kontexte und ihre Begriffe (1998). Lecture notes. →
69
[219] Kerkhoff, S., Pöschel, R., Schneider, F.M.: A short introduction to clones.
Electronic Notes in Theoretical Computer Science 303, 107–120 (2014).
doi:10.1016/j.entcs.2014.02.006. → 58
[220] Kipke, U., Wille, R.: Formale Begriffsanalyse erläutert an einem Wortfeld.
LDV-Forum 5, 31–36 (1987). → 113
[221] Kleitman, D.J.: Extremal properties of collections of subsets containing no
two sets and their union. Journal of Combinatorial Theory, Series A 20(3),
390–392 (1976). doi:10.1016/0097-3165(76)90037-6. → 69
[222] Klotz, U., Mann, A.: Begriffexploration. Diplomarbeit, TH Darmstadt
(1988). → 114
[223] Knecht, S., Wille, R.: Congruence lattices of finite lattices as concept lat-
tices. In [4] pp. 323–325 (1990). doi:10.1007/978-1-4899-2608-1_32. → 160
[224] Kollewe, W.: Representation of data by pseudoline arrangements. In [295]
pp. 113–122 (1993). → 113
[225] Kollewe, W., Skorsky, M., Vogt, F., Wille, R.: Toscana – ein Werkzeug zur
begrifflichen Analyse und Erkundung von Daten. In [419] pp. 267–288
(1994). → 113
[226] Koppen, M.: On finding the bidimension of a relation. Journal of
Mathematical Psychology 31(2), 155–178 (1987). doi:10.1016/0022-
2496(87)90013-7. → 281
[227] Korossy, K., et al.: Modeling knowledge as competence and perfor-
mance. In: D. Albert, J. Lukas (eds.) Knowledge spaces: Theories, em-
pirical research, and applications, pp. 103–132. Psychology Press (1999).
doi:10.4324/9781410602077-14. → 281
[228] Krajci, S.: Cluster based efficient generation of fuzzy concepts. Neural Net-
work World 13(5), 521–530 (2003). → 74
[229] Krantz, D., Luce, D., Suppes, P., Tversky, A.: Foundations of measurement,
Vol. I, II, 3. Academic Press, New York (1971–1990). → 71
[230] Kriegel, F.: Implications over probabilistic attributes. In [47] pp. 168–183
(2017). doi:10.1007/978-3-319-59271-8_11. → 74
References 349
[231] Kriegel, F.: Joining implications in formal contexts and inductive learning
in a Horn description logic. In [82] pp. 110–129 (2019). doi:10.1007/978-
3-030-21462-3_9. → 115
[232] Kriegel, F.: Constructing and Extending Description Logic Ontologies us-
ing Methods of Formal Concept Analysis: A Dissertation Summary. KI -
Künstliche Intelligenz 34, 399–403 (2020). doi:10.1007/s13218-020-00673-
8. → 333
[233] Kriegel, F.: Efficient axiomatization of OWL 2 EL ontologies from data by
means of Formal Concept Analysis. To appear (2024). → 333
[234] Kriegel, F., Borchmann, D.: Nextclosures: Parallel computation of the
canonical base with background knowledge. International Journal of Gen-
eral Systems 46(5), 490–510 (2017). doi:10.1080/03081079.2017.1349570.
→ 115
[235] Krolak-Schwerdt, S., Orlik, P., Ganter, B.: Tripat: a model for analyzing
three-mode binary data. In [55] pp. 298–307 (1994). doi:10.1007/978-
3-642-46808-7_27. → 73, 112
[236] Krötzsch, M.: Morphisms in Logic, Topology, and Formal Concept Analy-
sis. Master’s thesis, TU Darmstadt (2005). → 70, 309
[237] Krötzsch, M., Hitzler, P., Zhang, G.Q.: Morphisms in context. In: Concep-
tual Structures: Common Semantics for Sharing Knowledge. Proceedings
of the 13th ICCS., pp. 223–237. Springer (2005). doi:10.1007/11524564_15.
→ 157, 309
[238] Krötzsch, M., Malik, G.: The tensor product as a lattice of regular Galois
connections. In [284] pp. 89–104 (2006). doi:10.1007/11671404_6. → 162
[239] Krupka, M.: On factorization of concept lattices by incompatible toler-
ances. In [372] pp. 167–182 (2011). doi:10.1007/978-3-642-20514-9_14.
→ 160
[240] Krupka, M.: Basic theorem of fuzzy concept lattices revisited. Fuzzy Sets
and Systems 333, 54–70 (2018). doi:10.1016/j.fss.2017.04.007. → 73
[241] Krupka, M., Lastovicka, J.: Concept lattices of incomplete data. In [102]
pp. 180–194 (2012). doi:10.1007/978-3-642-29892-9_19. → 74
[242] Kryszkiewicz, M., Obiedkov, S. (eds.): Proceedings of the 7th CLA, vol.
CEUR 672 (2010). → 336, 345
[243] Kühn, R., Ries, M.: Methoden der Formalen Begriffsanalyse beim Aufbau
kooperativer Information. Diplomarbeit, TH Darmstadt (1988). → 113
[244] Kuznetsov, S.: Stability as an estimate of degree of substantiation of
hypotheses derived on the basis of operational similarity. Nauchno-
Tekhnicheskaya Informatsiya Ser. 2 24(12), 21–29 (1990). → 14
[245] Kuznetsov, S.O.: On computing the size of a lattice and related decision
problems. Order 18, 313–321 (2001). → 113, 249
[246] Kuznetsov, S.O.: Galois connections in data analysis: Contributions from
the soviet era and modern russian research. In [157] pp. 196–225 (2005).
doi:10.1007/11528784_11. → 20
350 References
[264] Luksch, P.: Distributive lattices freely generated by an ordered set of width
two. Discrete Mathematics 88(2-3), 249–258 (1991). doi:10.1016/0012-
365x(91)90013-r. → 71
[265] Luksch, P., Skorsky, M., Wille, R.: On drawing concept lattices with a com-
puter. In [166] (2001). → 113
[266] Luksch, P., Wille, R.: Formal concept analysis of paired comparisons. In
[53] pp. 567–576 (1987). → 214
[267] Luksch, P., Wille, R.: Substitution decomposition of concept lattices. In:
G. Eigenthaler, H.K. Kaiser, W.B. Müller, W. Nöbauer (eds.) Contributions
to general algebra, vol. 5, pp. 213–220. Hölder–Pichler–Tempsky, Wien
(1987). → 214
[268] Luksch, P., Wille, R.: A mathematical model for conceptual knowledge
systems. In [54] pp. 156–162 (1991). doi:10.1007/978-3-642-76307-6_21.
→ 332
[269] Luxenburger, M.: Partielle Implikationen und partielle Abhängigkeiten
zwischen Merkmalen. Diplomarbeit, TH Darmstadt (1988). → 113
[270] Luxenburger, M.: Implications partielles dans un contexte. Mathéma-
tiques informatique et sciences humaines 113, 35–55 (1991). → 113
[271] Luxenburger, M.: Implikationen, Abhängigkeiten und Galois Abbildun-
gen: Beiträge zur Formalen Begriffsanalyse. Ph.D. thesis, TH Darmstadt
(1993). → 113
[272] MacNeille, H.M.: Partially ordered sets. Transactions of the American
Mathematical Society 42(3), 416–460 (1937). doi:10.2307/1989739. → 71
[273] Maier, D.: The theory of relational databases, vol. 11. Computer science
press Rockville (1983). → 113
[274] Markowsky, G.: Some combinatorial aspects of lattice theory. Contract
14(67-A), 0298–0015 (1973). → 71
[275] Markowsky, G.: The factorization and representation of lattices. Trans-
actions of the American Mathematical Society 203, 185–200 (1975).
doi:10.2307/1997078. → 68
[276] Markowsky, G.: The representation of posets and lattices by sets. Algebra
Universalis 11(1), 173–192 (1980). → 281
[277] Marty, R.: Foliated semantic networks: concepts, facts, qualities. Com-
puters & mathematics with applications 23(6-9), 679–696 (1992).
doi:10.1016/0898-1221(92)90129-6. → 75
[278] Medina, J., Ojeda-Aciego, M., Ruiz-Calvino, J.: Formal concept analysis via
multi-adjoint concept lattices. Fuzzy sets and systems 160(2), 130–144
(2009). doi:10.1016/j.fss.2008.05.004. → 74
[279] Medina, R., Obiedkov, S. (eds.): Proceedings of the 6th ICFCA, vol.
LNAI 4933. Springer (2008). doi:10.1007/978-3-540-78137-0. → 340, 343,
346
[280] Mehrtens, H.: Die Entstehung der Verbandstheorie, vol. 6. Gerstenberg
(1979). → 20
[281] Meschke, C.: Concept approximations – approximative notions for con-
cept lattices. Ph.D. thesis, TU Dresden (2012). Qucosa Dresden, → 74
352 References
[299] Pensa, R.G., Boulicaut, J.F.: Towards fault-tolerant formal concept analysis.
In: Congress of the Italian Association for Artificial Intelligence, pp. 212–
223. Springer (2005). doi:10.1007/11558590_22. → 74
[300] Pierce, R.S.: Introduction to the theory of abstract algebras. Courier Cor-
poration (2014). doi:10.2307/2317361. → 214
[301] Poelmans, J., Ignatov, D.I., Kuznetsov, S.O., Dedene, G.: Fuzzy and rough
formal concept analysis: a survey. International Journal of General Sys-
tems 43(2), 105–134 (2014). doi:10.1080/03081079.2013.862377. → 74
[302] Pollandt, S.: Fuzzy-Begriffe: Formale Begriffsanalyse unscharfer Daten.
Springer-Verlag (2013). doi:10.1007/978-3-642-60460-7. → 357
[303] Pratt, V.: Chu spaces: Automata with quantum aspects. School on
Category Theory and Applications (Coimbra, 1999) 21, 39–100 (1999).
doi:10.1109/phycmp.1994.363682. → 71
[304] Prediger, S.: Kontextuelle Urteilslogik mit Begriffsgraphen: ein Beitrag zur
Restrukturierung der mathematischen Logik. Shaker (1998). → 333
[305] Prediger, S.: Mathematische Logik in der Wissensverar-
beitung. Mathematische Semesterberichte 47(2), 165–192 (2000).
doi:10.1007/s005910070002. → 333
[306] Prediger, S.: Terminologische Merkmalslogik in der Formalen Begriffs-
analyse. In [364] (2000). doi:10.1007/978-3-642-57217-3_5. → 331, 333
[307] Prediger, S., Stumme, G.: Theory-driven logical scaling. In: International
Workshop on Description Logics, vol. 22 (1999). → 333
[308] Priss, U.: The formalization of WordNet by methods of relational concept
analysis. In: WordNet: An Electronic Lexical Database and Some of its Ap-
plications. MIT press (1998). doi:10.7551/mitpress/7287.003.0013. → 334
[309] Priss, U., Polovina, S., Hill, R. (eds.): Proceedings of the 15th ICCS, vol.
LNAI 4604. Springer (2007). doi:10.1007/978-3-540-73681-3. → 352, 353
[310] Raney, G.N.: A subdirect-union representation for completely distribu-
tive complete lattices. Proceedings of the American Mathematical Society
4(4), 518–522 (1953). doi:10.2307/2032514. → 214, 280
[311] Reeg, S., Weiß, W.: Properties of finite lattices. Diplomarbeit, TH Darm-
stadt (1990). See also [145]. → 114, 279
[312] Reppe, H.: An FCA perspective on n-distributivity. In [309] pp. 255–268
(2007). doi:10.1007/978-3-540-73681-3_19. → 71, 280
[313] Reppe, H.: Three generalisations of lattice distributivity: An FCA perspec-
tive. Ph.D. thesis, TU Dresden (2010). Shaker, ISBN 978-3-8440-0037-5
→ 280
[314] Reuter, K.: Counting formulas for glued lattices. Order 1, 265–276 (1985).
doi:10.1007/bf00383603. → 160
[315] Reuter, K.: On the dimension of the cartesian product of relations and or-
ders. Order 6, 277–293 (1989). doi:10.1007/bf00563528. → 70, 281
[316] Reuter, K.: On the order dimension of convex polytopes. European Journal
of Combinatorics 11(1), 57–63 (1990). doi:10.1016/s0195-6698(13)80056-
x. → 281
354 References
[317] Reuter, K.: The jump number and the lattice of maximal antichains.
Discrete Mathematics 88(2-3), 289–307 (1991). doi:10.1016/0012-
365x(91)90016-u. → 71
[318] Reuter, K., Wille, R.: Complete congruence relations of concept lattices.
Acta Sci. Math. 51, 319–327 (1987). → 160
[319] Riguet, J.: Relations binaires, fermetures, correspondances de Galois.
Bulletin de la Société Mathématique de France 76, 114–155 (1948).
doi:10.24033/bsmf.1401. URL http://www.numdam.org/articles/10.
24033/bsmf.1401/. → 68
[320] Riguet, J.: Les relations de Ferrers. CR Acad. Sci. Paris 232(1729), 116
(1951). → 281
[321] Roberts, F.S.: Measurement theory. Addison-Wesley, Reading, Mass.
(1979). → 99
[322] Rouane-Hacene, M., Huchard, M., Napoli, A., Valtchev, P.: A proposal for
combining formal concept analysis and description logics for mining re-
lational data. In [251] pp. 51–65 (2007). doi:10.1007/978-3-540-70901-5_4.
→ 334
[323] Rudolph, S.: Relational Exploration - Combining Description Logics and
Formal Concept Analysis for Knowledge Specification. Universitätsverlag
Karlsruhe (2006). → 326, 331, 333
[324] Rudolph, S.: Some notes on managing closure operators. In [102] pp. 278–
291 (2012). doi:10.1007/978-3-642-29892-9_25. → 21
[325] Rudolph, S.: On the succinctness of closure operator representations.
In [173] pp. 15–36 (2014). doi:10.1007/978-3-319-07248-7_2. → 21
[326] Rudolph, S., Săcărea, C., Troancă, D.: Towards a navigation paradigm for
triadic concepts. In: Formal Concept Analysis: 13th International Con-
ference, ICFCA 2015, Nerja, Spain, June 23-26, 2015, Proceedings 13, pp.
252–267. Springer (2015). doi:10.1007/978-3-319-19545-2_16. → 73
[327] Rusch, A., Wille, R.: Knowledge spaces and formal concept analysis. In
[56] p. 427ff. (1986). doi:10.1007/978-3-642-80098-6_36. → 113
[328] Ryssel, U., Distel, F., Borchmann, D.: Fast algorithms for implication bases
and attribute exploration using proper premises. Annals of Mathemat-
ics and Artificial Intelligence pp. 101–113 (2014). doi:10.1007/s10472-013-
9355-9. → 115
[329] Săcărea, C.: Towards a theory of contextual topology. Shaker Verlag
(2001). → 75
[330] Sakurai, T.: On formal concepts of random formal contexts. Information
Sciences 578, 615–620 (2021). doi:10.1016/j.ins.2021.07.065. → 69
[331] Schader, M. (ed.): Analysing and modelling data and knowledge.
Springer–Verlag (1992). → 356
[332] Schaffert, S.: Zerlegungssätze für symmetrische Kontexte. Master’s thesis,
TH Darmstadt (1985). → 70, 71
[333] Scheich, P., Skorsky, M., Vogt, F., Wachter, C., Wille, R.: Conceptual data
systems. In [295] pp. 72–84 (1993). doi:10.1007/978-3-642-50974-2_8.
→ 71
References 355
[334] Schlemmer, T.: Annotating lattice orbifolds with minimal acting automor-
phisms. In: CLA’12, vol. CEUR-WS 972, pp. 57–68 (2012). → 308
[335] Schmid, J.: Bialgebraic contexts for distributive lattices–revisited. In [149]
pp. 403–407 (2005). doi:10.1007/978-3-540-32262-7_28. → 357
[336] Schmidt, B.: Ein Zusammenhang zwischen Graphentheorie und Verband-
stheorie. Diplomarbeit, TH Darmstadt (1982). → 71
[337] Schmidt, J.: Zur Kennzeichnung der Dedekind-MacNeilleschen Hülle
einer geordneten Menge. Archiv der Mathematik 7(4), 241–249 (1956).
doi:10.1007/bf01900297. → 68, 71, 281
[338] Schmidt, R.: Verbandstheoretische Analyse von Liniendiagrammen.
Diplomarbeit, TH Darmstadt (1988). → 280
[339] Schütt, D.: Abschätzungen für die Anzahl der Begriffe von Kontexten.
Diplomarbeit, TH Darmstadt (1987). → 112
[340] Sertkaya, B.: Formal concept analysis methods for desciption logics. Ph.D.
thesis, TU Dresden (2007). Qucosa Dresden → 331, 333
[341] Sertkaya, B., Oğuztüzün, H.: Proof of the basic theorem on concept lattices
in Isabelle/HOL. In: International Symposium on Computer and Informa-
tion Sciences, pp. 976–985. Springer (2004). doi:10.1007/978-3-540-30182-
0_98. → 68
[342] Shmuely, Z.: The structure of Galois connections. Pacific Journal of Math-
ematics 54(2), 209–225 (1974). doi:10.2140/pjm.1974.54.209. → 214
[343] Skorsky, M.: How to draw a concept lattice with parallelograms. In [401]
(1989). → 113, 280
[344] Skorsky, M.: Endliche Verbände: Diagramme und Eigenschaften. Ph.D.
thesis, TH Darmstadt (1992). → 112, 113, 279
[345] Soldano, H., Ventos, V.: Abstract concept lattices. In [372] pp. 235–250
(2011). doi:10.1007/978-3-642-20514-9_18. → 74
[346] Sowa, J.F.: Conceptual graphs summary. Conceptual Structures: current
research and practice 3, 66 (1992). → 333
[347] Spangenberg, N., Wolff, K.E.: Comparison of biplot analysis and formal
concept analysis in the case of a repertory grid. In: Classification, Data
Analysis, and Knowledge Organization: Models and Methods with Appli-
cations, pp. 104–112. Springer (1991). doi:10.1007/978-3-642-76307-6_15.
→ 72
[348] Stahl, J., Wille, R.: Preconcepts and set representations of contexts. In [166]
pp. 431–438 (1986). → 281, 332
[349] Stanley, R.P.: Enumerative combinatorics volume 1 second edi-
tion. Cambridge studies in advanced mathematics (2011).
doi:10.1017/cbo9781139058520. → 14
[350] Stephan, J.: Substitution decomposition of lattices. Ph.D. thesis, TH Darm-
stadt (1991). → 214
[351] Stephan, J.: Substitution products of lattices. In [106] pp. 321–336 (1991).
→ 214
[352] Stern, M.: Semimodular lattices: theory and applications, vol. 73. Cam-
bridge University Press (1999). doi:10.1017/cbo9780511665578. → 280
356 References
[353] Stöhr, B., Wille, R.: Formal concept analysis of data with tolerances. In:
Analyzing and Modeling Data and Knowledge: Proceedings, pp. 117–127.
Springer (1991). doi:10.1007/978-3-642-46757-8_13. → 114
[354] Strahringer, S.: Direct products of convex-ordinal scales. Order 11, 361–383
(1994). doi:10.1007/bf01108768. → 72, 247
[355] Strahringer, S.: Dimensionality of ordinal structures. Discrete Mathemat-
ics 144(1-3), 97–117 (1995). doi:10.1016/0012-365x(94)00289-u. → 72
[356] Strahringer, S., Wille, R.: Convexity in ordinal data. In [54] pp. 113–120
(1991). doi:10.1007/978-3-642-76307-6_16. → 72
[357] Strahringer, S., Wille, R.: Towards a structure theory for ordinal data. In
[331] pp. 129–139 (1992). doi:10.1007/978-3-642-46757-8_14. → 72
[358] Strahringer, S., Wille, R.: Conceptual clustering via convex-ordinal struc-
tures. In [295] pp. 85–98 (1993). doi:10.1007/978-3-642-50974-2_9. → 72
[359] Stumme, G.: Attribute exploration with background implications and ex-
ceptions. In [56] pp. 457–469 (1996). doi:10.1007/978-3-642-80098-6_39.
→ 114
[360] Stumme, G.: The concept classification of a terminology extended by con-
junction and disjunction. In: PRICAI’96: Topics in Artificial Intelligence:
4th Pacific Rim International Conference on Artificial Intelligence Cairns,
Australia, August 26–30, 1996 Proceedings 4, pp. 121–131. Springer (1996).
doi:10.1007/3-540-61532-6_11. → 333
[361] Stumme, G.: Distributive concept exploration—a knowledge acquisition
tool in formal concept analysis. In: Annual Conference on Artificial Intel-
ligence, pp. 117–128. Springer (1998). doi:10.1007/bfb0095433. → 114,
247
[362] Stumme, G., Taouil, R., Bastide, Y., Pasquier, N., Lakhal, L.: Computing
iceberg concept lattices with titanic. Data & knowledge engineering 42(2),
189–222 (2002). doi:10.1016/s0169-023x(02)00057-5. → 279
[363] Stumme, G., Wille, R.: A geometrical heuristic for drawing concept lattices.
In: R. Tamassia, I. Tollis (eds.) Graph Drawing: DIMACS GD’84, pp. 452–
459. Springer (1995). doi:10.1007/3-540-58950-3_399. → 113
[364] Stumme, G., Wille, R.: Begriffliche Wissensverarbeitung: Methoden und
Anwendungen. Springer-Verlag (2000). doi:10.1007/978-3-642-57217-3.
→ 343, 353
[365] Szathmary, L., Valtchev, P., Napoli, A., Godin, R., Boc, A., Makarenkov, V.:
A fast compound algorithm for mining generators, closed itemsets, and
computing links between equivalence classes. Ann. Math. Artif. Intell.
70(1-2), 81–105 (2014). doi:10.1007/s10472-013-9372-8. → 115
[366] Takàcs, V.: Two applications of Galois graphs in pedagogical research
(Feb. 8, 1984). Manuscript of a lecture given at TH Darmstadt. → 68
[367] Tamari, D.: The algebra of bracketings and their enumeration. Nieuw
Arch. Wisk 3(10), 131–146 (1962). → 235, 247
[368] Teo, S.: Representing finite lattices as complete congruence lattices of com-
plete lattices. Ann. Univ. Sci. Budapest. Eötvös Sect. Math. 33, 177–182
(1990). → 160
References 357
[404] Wille, R.: The skeletons of free distributive lattices. Discrete Mathematics
88(2-3), 309–320 (1991). doi:10.1016/0012-365x(91)90017-v. → 71
[405] Wille, R.: Tensor products of complete lattices as closure systems. In [106]
pp. 381–385 (1991). → 214
[406] Wille, R.: Begriffliche Datensysteme als Werkzeug der Wissenskommu-
nikation. In: H.H. Zimmermann, H.D. Luckhardt, A. Schulz (eds.) Men-
sch und Maschine – Informationelle Schnittstellen der Kommunikation,
pp. 63–73. Universitätsverlag Konstanz (1992). → 113
[407] Wille, R.: Concept lattices and conceptual knowledge systems. Com-
puters & mathematics with applications 23(6-9), 493–515 (1992).
doi:10.1016/0898-1221(92)90120-7. → 71, 113
[408] Wille, R.: Plädoyer für eine philosophische Grundlegung der Begrifflichen
Wissensverarbeitung. In [419] pp. 11–25 (1994). → 68
[409] Wille, R.: The basic theorem of triadic concept analysis. Order 12, 149–158
(1995). doi:10.1007/bf01108624. → 73
[410] Wille, R.: Begriffsdenken: Von der griechischen Philosophie bis zur Kün-
stlichen Intelligenz heute. Dilthey-Kastanie pp. 11–25 (1995). → 68
[411] Wille, R.: Restructuring mathematical logic: an approach based on Peirce’s
pragmatism. In: P.A. Aldo Ursini (ed.) Logic and algebra, pp. 267–281.
Marcel Dekker, New York (1996). doi:10.1201/9780203748671-12. → 75
[412] Wille, R.: Boolean concept logic. In [152] pp. 317–331 (2000).
doi:10.1007/10722280_22. → 311, 322, 324, 332
[413] Wille, R.: Boolean judgment logic. In [92] pp. 115–128 (2001).
doi:10.1007/3-540-44583-8_9. → 311
[414] Wille, R.: Contextual logic summary. Math. Preprint 2098, TH Darmstadt
(2001). → 333
[415] Wille, R.: Existential concept graphs of power context families. In: Interna-
tional Conference on Conceptual Structures, pp. 382–395. Springer (2002).
doi:10.1007/3-540-45483-7_29. → 334
[416] Wille, R.: Truncated distributive lattices: Conceptual structures
of simple-implicational theories. Order 20(3), 229–238 (2003).
doi:10.1023/b:orde.0000026494.22248.85. → 332
[417] Wille, R.: Preconcept algebras and generalized double Boolean algebras.
In [114] pp. 1–13 (2004). doi:10.1007/978-3-540-24651-0_1. → 318, 321,
332
[418] Wille, R.: Formal concept analysis and contextual logic. In: P. Hitzler,
H. Schärfe (eds.) Conceptual Structures in Practice, pp. 155–192. Chap-
man and Hall/CRC (2016). doi:10.1201/9781420060638-16. → vii
[419] Wille, R., Zickwolff, M.: Begriffliche Wissensverarbeitung: Grundfragen
und Aufgaben. B.I. Wissenschaftsverlag (1994). → 348, 359
[420] Wille, U.: Eine Axiomatisierung bilinearer Kontexte. Mitt. Math. Sem.
Gießen 200, 71–112 (1991). Diplomarbeit, → 75
[421] Wille, U.: Geometric representation of ordinal contexts. Ph.D. thesis, Uni-
versität Giessen (1995). → 75
360 References
[422] Wolff, K.E.: Concepts in fuzzy scaling theory: order and granularity. Fuzzy
Sets and Systems 132(1), 63–75 (2002). doi:10.1016/s0165-0114(02)00106-
9. → 73
[423] Wolff, K.E.: Ordnung, Wille und Begriff. Geschichten aus dem Ernst
Schröder Zentrum (2003). www.ernst-schroeder-zentrum.de → vii
[424] Wolff, K.E.: Applications of temporal conceptual semantic systems. In:
International Conference on Knowledge Processing in Practice, pp. 59–78.
Springer (2007). doi:10.1007/978-3-662-55433-3_11. → 75
[425] Xia, W.: Morphismen als formale Begriffe: Darstellung und Erzeugung.
Ph.D. thesis, TH Darmstadt (1993). → 161, 296, 308
[426] Yanagimoto, T., Okamoto, M.: Partial orderings of permutations and
monotonicity of a rank correlation statistic. Annals of the Institute of Sta-
tistical Mathematics 21, 489–506 (1969). doi:10.1007/bf02532273. → 55
[427] Zhi, H., Li, J.: Granule description of incomplete data: a cogni-
tive viewpoint. Cognitive Computation 14(6), 2108–2119 (2022).
doi:10.1007/s12559-021-09918-6. → 74
[428] Zickwolff, M.: Darstellung symmetrischer Strukturen durch Transversale.
In [106] pp. 391–403 (1991). → 308
[429] Zickwolff, M.: Rule exploration: first order logic in formal concept analy-
sis. Ph.D. thesis, TH Darmstadt (1991). → 114, 308
[430] Zschalig, C.: An fdp-algorithm for drawing lattices. In: CLA’07,
vol. CEUR-WS 331 (2007). → 113
Formal contexts and concept lattices in this
book
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 361
B. Ganter, R Wille, Formal Concept Analysis, https://doi.org/10.1007/978-3-031-63422-2
Index
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 363
B. Ganter, R Wille, Formal Concept Analysis, https://doi.org/10.1007/978-3-031-63422-2
364 Index
non-redundant, 96 Pol-Inv, 58
polarity, 52
object, 23 polarity lattice, 52
-reduced, 31 polyhedron, 12
concept, 29 Pontryagin duality, 57
intent, 28 positioning rule, 88
reducible, 31 power context families, 334
one-valued context, 59 powerset, 2
opposition of formal concepts, 321 preconcept, 276, 317
orbit-maximal, 285 preconcept algebra, 321
Order, 279 preconcepts
order, 2 ordering of, 318
-embedding, 3 premise
-isomorphism, 3 of an implication, 93
-preserving map, 3 principal filter, 3
-reversing map, 4 principal ideal, 3
of formal concepts, 25 product, 4, see also direct product
relation, 2 product measure, 302
as a lattice, 44 projective plane, 57
order dimension, 270 proper premise, 95
order extension, 5 of an attribute, 96
order filter, 48 protoconcept, 317
order folding, 287 pseudo-complement, 307
order ideal, 48 pseudo-intent, 97
order relation, 2
formal context of, 44 Q-atlas, 171
ordered set, 2
ordinal factor analysis, 281 random
ordinal lattice, 65 closure system, 70
ordinal scale formal context, 70
general, 46 rank function, 258
one-dimensional, 64 reduced, 31
ordinally dependent, 110 labelling, 29
ortholattice, 53 regular
orthomodular, 57 clause set, 314
overlapping neighborhoods, 173 relation
k-ary, 329
P -context, 219 binary, 1
P -fusion, 221 relational concept analysis, 334
P -lattice, 218 residual map, 20
P -product, 219 residuated map, 20
partial implications, 113 respect
partial order, 2 an implication, 93
pattern structures, 74 retract, 161
plain scaling, 60 rigging of a substitution sum, 186
Index 369
unfolding, 290
union of formal contexts, 225
unit element, 5
upper bound, 5
upper neighbor, 2
valence
of a lattice element, 307
of an object, 306
vertical sum, 42
view
conceptual, 66
contextual, 66
zero element, 5