Research Article A Discrete Dynamical Model of Signed Partitions

advertisement
Hindawi Publishing Corporation
Journal of Applied Mathematics
Volume 2013, Article ID 973501, 10 pages
http://dx.doi.org/10.1155/2013/973501
Research Article
A Discrete Dynamical Model of Signed Partitions
G. Chiaselotti, G. Marino, P. A. Oliverio, and D. Petrassi
Dipartimento di Matematica, UniversitaΜ€ della Calabria, Via Pietro Bucci, Cubo 30B, 87036 Arcavacata di Rende, Italy
Correspondence should be addressed to G. Marino; gmarino@unical.it
Received 28 November 2012; Accepted 19 February 2013
Academic Editor: Luigi Muglia
Copyright © 2013 G. Chiaselotti et al. This is an open access article distributed under the Creative Commons Attribution License,
which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
We use a discrete dynamical model with three evolution rules in order to analyze the structure of a partially ordered set of signed
integer partitions whose main properties are actually not known. This model is related to the study of some extremal combinatorial
sum problems.
1. Introduction
A dynamical system is by definition a system whose state
changes with time 𝑑. We have a discrete dynamical system
when 𝑑 is an integer or a natural number, and the elements of
the system can be obtained in the form 𝑒𝑑+1 = 𝐹(𝑒𝑑 ), where 𝐹
is some global function which describes the evolution rule of
the system (see [1]). In a series of very recent works, the theory
of discrete dynamical systems has been applied in several
contexts. In [2], the theory of discrete dynamical systems
is applied in order to analyze some models of concurrent
computing systems. In [3], a dynamical model of parallel
computation on bi-infinite time scale with an approach
similar to two-sided symbolic dynamics is constructed. In
[4], the authors analyze the orbit structure of parallel discrete
dynamical systems over directed dependency graphs, with
Boolean functions as global functions. In [5], the authors
extend the manner of defining the evolution update of
discrete dynamical systems on Boolean functions, without
limiting the local functions to being dependent restrictions
of a global one. Finally, in [6] is given a complete characterization of the orbit structure of parallel discrete dynamical
systems with maxterm and minterm Boolean functions as
global functions. In [7], the authors have introduced the poset
(𝑆(𝑛, 𝑑, π‘Ÿ), βŠ‘) of all the signed integer partitions π‘Ÿ ≥ π‘Žπ‘Ÿ ≥ ⋅ ⋅ ⋅ ≥
π‘Ž1 ≥ 0 ≥ 𝑏1 ≥ ⋅ ⋅ ⋅ ≥ 𝑏𝑛−π‘Ÿ ≥ −(𝑛 − π‘Ÿ), where the positive parts
and the negative parts are all distinct between them and the
number of nonzero parts is exactly 𝑑. In this case, the partial
order βŠ‘ is that on the components. The concept of signed
integer partition has been introduced in [8] and studied in
[9] from an arithmetical point of view. In [7, 10], it is shown
that the structure of the lattice 𝑆(𝑛, 𝑑, π‘Ÿ) is strictly related to
the study of some combinatorial extremal sum problems. In
this paper, we study some particular type of signed partitions
(specifically, of the signed partitions of 𝑆(𝑛, 𝑑, π‘Ÿ)) by adopting
the interesting point of view of the discrete dynamical
models. In particular, we analyze the lattice 𝑆(𝑛, 𝑑, π‘Ÿ) as a
discrete dynamical model with three local evolution rules
whose dynamics is studied in a sequential mode. The way
to study a lattice of classical partitions as a discrete dynamic
model having some particular evolution rules is implicit in
[11], where Brylawski proposed a dynamical approach to
study the lattice 𝐿 𝐡 (𝑛) of all the partitions of the fixed positive
integer 𝑛 with the dominance ordering. In such a context,
a configuration of the system is represented by an ordered
partition of an integer 𝑛, that is, a decreasing sequence π‘Ž =
(π‘Ž1 , . . . , π‘Žπ‘› ) having sum 𝑛, whose blocks of the corresponding
Young diagram are considered as mobile sand grain, and the
movement of a sand grain, respects the following rules.
Rule 1 (vertical rule). One grain can move from a column
to the next column if the difference of height of these two
columns is greater than or equal to 2.
Rule 2 (horizontal rule). If a column containing 𝑝 + 1 grains
is followed by a sequence of columns containing 𝑝 grains and
then one column containing 𝑝 − 1 grains, then one grain of
the first column can slip to the last column.
In this paper, we prove that the covering relation in the
lattice 𝑆(𝑛, 𝑑, π‘Ÿ) is uniquely determined by three evolution
rules of our discrete dynamical model. We articulate the proof
2
Journal of Applied Mathematics
of this result by contradiction by using only the first-order
logic. This method of proof is very laborious; however, its
advantage is that it can be easily implemented on a proof
assistant checker based on the first-order logic (e.g., Mizar).
The paper is articulated as follows. In Section 2, we recall
some basic definitions and preliminary results, for example,
the definition of 𝑆(𝑛, 𝑑, π‘Ÿ) and some of its properties. In
Section 3, we explain how to see the signed partitions of
𝑆(𝑛, 𝑑, π‘Ÿ) as configurations of our discrete dynamical model
and we also describe its evolution rules. In Section 4, we
assume by contradiction that there is a generic element 𝑀󸀠
of 𝑆(𝑛, 𝑑, π‘Ÿ) that covers another generic element 𝑀 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ)
but 𝑀󸀠 is not generated by 𝑀 with none of our evolution rules.
From this assumption we deduce several propositions and
conditions that we use in order to prove our principal result,
that is, Theorem 4.
2. Definitions and Preliminary Results
(I2) if π‘₯, 𝑦 ∈ 𝑋 and if π‘₯ ≤ 𝑦, then 𝑦𝑐 ≤ π‘₯𝑐 .
The map 𝑐 is called complementation of 𝑋 and π‘₯𝑐 the
complement of π‘₯. Let us observe that if 𝑋 is an involution
poset, by (𝐼1) it follows that 𝑐 is bijective and by (𝐼1) and
(𝐼2) it holds that if π‘₯, 𝑦 ∈ 𝑋 are such that π‘₯ < 𝑦, then
𝑦𝑐 < π‘₯𝑐 . If (𝑋, ≤, 𝑐) is an involution poset and if 𝑍 ⊆ 𝑋, we
will set 𝑍𝑐 = {𝑧𝑐 : 𝑧 ∈ 𝑍}. We note that if 𝑋 is an involution
poset, then 𝑋 is a self-dual poset because from (𝐼1) and (𝐼2) it
follows that if π‘₯, 𝑦 ∈ 𝑋, we have that π‘₯ ≤ 𝑦, if and only if 𝑦𝑐 ≤
π‘₯𝑐 , and this is equivalent to say that the complementation is
an isomorphism between 𝑋 and its dual poset 𝑋∗ . In [7], it
has been shown that (𝑆(𝑛, π‘Ÿ), βŠ‘) is an involution poset and its
complementation map 𝑐 is the following:
𝑐
(π‘Žπ‘˜ ⋅ ⋅ ⋅ π‘Ž1 0 ⋅ ⋅ ⋅ 0 | 0 ⋅ ⋅ ⋅ 0 𝑏1 ⋅ ⋅ ⋅ 𝑏𝑙 )
σΈ€ 
σΈ€ 
= π‘Žπ‘Ÿ−π‘˜
⋅ ⋅ ⋅ π‘Ž1σΈ€  0 ⋅ ⋅ ⋅ 0 | 0 ⋅ ⋅ ⋅ 0 𝑏1σΈ€  ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ−𝑙
,
If (𝑋, ≤) is a poset and π‘₯, 𝑦 ∈ 𝑋, we write 𝑦 β‹— π‘₯ (or π‘₯ β‹–
𝑦) if 𝑦 covers π‘₯. Now we briefly recall the definition of the
lattice 𝑆(𝑛, π‘Ÿ) that we have introduced in [7] in a more formal
context. Let 𝑛 and π‘Ÿ be two nonnegative integers such that
π‘Ÿ ≤ 𝑛.
We call (𝑛, π‘Ÿ)-string a 𝑛-tuplet of integers
π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 | 𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ ,
(I1) (π‘₯𝑐 )𝑐 = π‘₯, for all π‘₯ ∈ 𝑋;
(1)
such that
(i) π‘Ž1 , . . . , π‘Žπ‘Ÿ ∈ {1, . . . , π‘Ÿ, 0};
(ii) 𝑏1 , . . . , 𝑏𝑛−π‘Ÿ ∈ {−1, . . . , −(𝑛 − π‘Ÿ), 0};
(iii) π‘Žπ‘Ÿ ≥ ⋅ ⋅ ⋅ ≥ π‘Ž1 ≥ 0 ≥ 𝑏1 ≥ ⋅ ⋅ ⋅ ≥ 𝑏𝑛−π‘Ÿ ;
(iv) the unique element in (1) which can be repeated is 0.
If 𝑀 is a (𝑛, π‘Ÿ)-string, we call parts of 𝑀 the integers
π‘Žπ‘Ÿ , . . . , π‘Ž1 , 𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ , nonnegative parts of 𝑀 the integers
π‘Žπ‘Ÿ , . . . , π‘Ž1 , and nonpositive parts of 𝑀 the integers 𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ .
We set 𝑀+ = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 | and 𝑀− = |𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ . In some cases,
we do not distinguish between nonnegative and nonpositive
parts of 𝑀, and we write more simply 𝑀 = 𝑙1 ⋅ ⋅ ⋅ 𝑙𝑛 instead of
𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 |𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ . We also denote by |𝑀|> the number
of parts of 𝑀 that are strictly positive, with |𝑀|< the number
of parts of 𝑀 that are strictly negative, and we set ‖𝑀‖ =
|𝑀|> + |𝑀|< . 𝑆(𝑛, π‘Ÿ) is the set of all the (𝑛, π‘Ÿ)-strings. If 𝑀 =
σΈ€ 
are two (𝑛, π‘Ÿ)π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 |𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ and 𝑀󸀠 = π‘Žπ‘ŸσΈ€  ⋅ ⋅ ⋅ π‘Ž1σΈ€  |𝑏1σΈ€  ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ
σΈ€ 
σΈ€ 
strings, we set 𝑀+ = 𝑀+ if π‘Žπ‘– = π‘Žπ‘– for all 𝑖 = π‘Ÿ, . . . , 1, 𝑀− = 𝑀−σΈ€ 
if 𝑏𝑗󸀠 = 𝑏𝑗 for all 𝑗 = 1, . . . , 𝑛 − π‘Ÿ, and 𝑀 = 𝑀󸀠 if 𝑀+ = 𝑀+σΈ€ 
and 𝑀− = 𝑀−σΈ€  . On 𝑆(𝑛, π‘Ÿ), we consider the partial order on the
components that we denote by βŠ‘. To simplify the notations,
in all the numerical examples the integers on the right of
the vertical bar | will be written without minus sign. Since
(𝑆(𝑛, π‘Ÿ), βŠ‘) is a finite distributive lattice, it is also graded, with
minimal element 0 ⋅ ⋅ ⋅ 0|12 ⋅ ⋅ ⋅ (𝑛 − π‘Ÿ) and maximal element
π‘Ÿ(π‘Ÿ − 1) ⋅ ⋅ ⋅ 21|0 ⋅ ⋅ ⋅ 0.
We recall now the concept of involution poset (see [12, 13]
for some recent studies on such class of posets). An involution
poset (IP) is a poset (𝑋, ≤, 𝑐) with a unary operation 𝑐 : π‘₯ ∈
𝑋 󳨃→ π‘₯𝑐 ∈ 𝑋, such that
(2)
σΈ€ 
where {π‘Ž1σΈ€  , . . . , π‘Žπ‘Ÿ−π‘˜
} is the usual complement of {π‘Ž1 , . . . , π‘Žπ‘˜ }
σΈ€ 
} is the usual complement of
in {1, . . . , π‘Ÿ}, and {𝑏1σΈ€  , . . . , 𝑏𝑛−π‘Ÿ−𝑙
{𝑏1 , . . . , 𝑏𝑙 } in {−1, . . . , −(𝑛 − π‘Ÿ)} (e.g., in 𝑆(7, 4), we have that
(4310|001)𝑐 = 2000|023). If 𝑑 is an integer such that 0 ≤
𝑑 ≤ 𝑛, we set now 𝑆(𝑛, 𝑑, π‘Ÿ) = {𝑀 ∈ 𝑆(𝑛, π‘Ÿ) : ‖𝑀‖ = 𝑑}. Its
easy to see that (𝑆(𝑛, 𝑑, π‘Ÿ), βŠ‘) is a sublattice of (𝑆(𝑛, π‘Ÿ), βŠ‘) and
obviously |𝑆(𝑛, 𝑑, π‘Ÿ)| = ( 𝑑𝑛 ).
We call signed Young diagrams (briefly SYD) an ordered
couple 𝐷 = 𝐷1 : 𝐷2 , where 𝐷1 is a Young diagram of an
integer partition built with decreasing columns instead with
decreasing rows, and 𝐷2 is a Young diagram of an integer
partition built with increasing columns. In the sequel, we will
call pile a column of a Young diagram and grain a square of a
pile. For example,
(3)
𝐷=
:
is determined by the partitions 4331 and 113. If 𝑀 =
π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 |𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ ∈ 𝑆(𝑛, π‘Ÿ), we denote by 𝐷(𝑀) the signed
Young diagram 𝐷(𝑀) = 𝐷+ (𝑀) : 𝐷− (𝑀), where 𝐷+ (𝑀) is
the Young diagram of the partition (π‘Žπ‘Ÿ , . . . , π‘Ž1 ) and 𝐷− (𝑀) is
the Young diagram of the partition (−𝑏1 , . . . , −𝑏𝑛−π‘Ÿ ). Now, if
𝐷 = 𝐷1 : 𝐷2 is a SYD, where (π‘Žπ‘‘ , . . . , π‘Ž1 ) is the decreasing
partition having 𝐷1 as Young diagram and ((−𝑏1 ), . . . , (−𝑏𝑠 ))
is the increasing partition having 𝐷2 as Young diagram, we
set 𝑝(𝐷) = π‘Žπ‘‘ ⋅ ⋅ ⋅ π‘Ž1 |(−𝑏1 ) ⋅ ⋅ ⋅ (−𝑏𝑠 ) and we call 𝑝(𝐷) the signed
partition of 𝐷. We say that a signed Young diagram 𝐷 is a
(𝑛, π‘Ÿ)-SYD if it results that 𝑝(𝐷) ∈ 𝑆(𝑛, π‘Ÿ) and we denote with
π‘Œπ‘†(𝑛, π‘Ÿ) the set of all the (𝑛, π‘Ÿ)-SYD. It is clear that the map πœ™ :
𝑀 󳨃→ 𝐷(𝑀) is a bijection between 𝑆(𝑛, π‘Ÿ) and π‘Œπ‘†(𝑛, π‘Ÿ), whose
inverse function is the map πœ“ : 𝐷 󳨃→ 𝑝(𝐷); therefore, in the
sequel, we will identify the elements of 𝑆(𝑛, π‘Ÿ) with the (𝑛, π‘Ÿ)SYDs of π‘Œπ‘†(𝑛, π‘Ÿ). If 𝐴 and 𝐡 are two usual Young diagrams,
we set 𝐴 ⊆ 𝐡 if 𝐴 is a subdiagram of 𝐡. If 𝐷 = 𝐷1 : 𝐷2 and
𝐷󸀠 = 𝐷1σΈ€  : 𝐷2σΈ€  are two SYDs, we set
𝐷 βŠ‘ 𝐷󸀠 ⇐⇒ 𝐷1 ⊆ 𝐷1σΈ€  ,
𝐷2 ⊇ 𝐷2σΈ€  .
(4)
Journal of Applied Mathematics
3
Therefore, it is clear that we can identify (𝑆(𝑛, π‘Ÿ), βŠ‘) with
(π‘Œπ‘†(𝑛, π‘Ÿ), βŠ‘). Now we set π‘Œπ‘†(𝑛, 𝑑, π‘Ÿ) = {𝐷(𝑀) : 𝑀 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ)}.
Obviously we can identify (π‘Œπ‘†(𝑛, 𝑑, π‘Ÿ), βŠ‘) with (𝑆(𝑛, 𝑑, π‘Ÿ), βŠ‘).
If (𝐻) is a statement, in the sequel we denote by ¬(𝐻) the
negation of (𝐻).
Rule 3. One grain must be deleted from the 𝑗th-minus pile if
𝑀 has a minus cliff at 𝑗:
βˆ™
:
3. Description of the Evolution Rules
In our dynamic model, the set of all configurations is exactly
π‘Œπ‘†(𝑛, 𝑑, π‘Ÿ). Our goal is to define some rules of evolution
that starting from the minimum of π‘Œπ‘†(𝑛, 𝑑, π‘Ÿ) allow us to
reconstruct the Hasse diagram of π‘Œπ‘†(𝑛, 𝑑, π‘Ÿ) (and therefore
to determine the covering relations in 𝑆(𝑛, 𝑑, π‘Ÿ)).
Let 𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 |𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ ∈ 𝑆(𝑛, 𝑑, π‘Ÿ) and 𝐷 = 𝐷(𝑀) =
𝐷+ (𝑀) : 𝐷− (𝑀). If 1 ≤ 𝑖 ≤ π‘Ÿ, we call 𝑖th-plus pile
of 𝐷, and we denote it by 𝐢+ (𝑖, 𝐷), the 𝑖th pile of 𝐷+ (𝑀)
that corresponds to part π‘Žπ‘– of 𝑀 and, if 1 ≤ 𝑗 ≤ 𝑛 − π‘Ÿ,
we call 𝑗th-minus pile of 𝐷, and we denote it by 𝐢− (𝑗, 𝐷),
the 𝑗th pile of 𝐷− (𝑀) that corresponds to part 𝑏𝑗 of 𝑀.
We call the pile 𝐢+ (𝑖, 𝐷) 𝑝𝑙𝑒𝑠 π‘ π‘–π‘›π‘”π‘™π‘’π‘‘π‘œπ‘› 𝑝𝑖𝑙𝑒 if π‘Žπ‘– = 1 and
𝐢− (𝑗, 𝐷) π‘šπ‘–π‘›π‘’π‘  π‘ π‘–π‘›π‘”π‘™π‘’π‘‘π‘œπ‘› 𝑝𝑖𝑙𝑒 if 𝑏𝑗 = 1. Let us note that if
there exists a plus singleton pile, then it is necessarily unique,
analogously for a minus singleton pile. If 1 < 𝑖 ≤ π‘Ÿ, we set
Δ+𝑖 (𝑀) = π‘Žπ‘– − π‘Žπ‘–−1 and we call Δ+𝑖 (𝑀) the plus height difference
of 𝑀 in 𝑖. If 1 < 𝑗 ≤ 𝑛 − π‘Ÿ, we set Δ−𝑗 (𝑀) = |𝑏𝑗 | − |𝑏𝑗−1 | and we
call Δ−𝑗 (𝑀) the minus height difference of 𝑀 in 𝑗. If 1 < 𝑖 ≤ π‘Ÿ,
we say that 𝑀 has a plus cliff at 𝑖 if Δ+𝑖 (𝑀) ≥ 2. If 1 < 𝑗 ≤ 𝑛 − π‘Ÿ,
we say that 𝑀 has a minus cliff at 𝑗 if Δ−𝑗 (𝑀) ≥ 2.
Remark 1. When we apply the following rules to one element
𝑀 ∈ 𝑆(𝑛, π‘Ÿ), we impose that there is an “invisible” extra pile
in the imaginary place (π‘Ÿ + 1) of 𝐷+ (𝑀) having exactly π‘Ÿ + 1
grains. This is only a formal trick for decreas the number of
rules necessary for our model; therefore, the (π‘Ÿ+1)th pile must
be not considered as a part of 𝑀.
3.1. Evolution Rules.
Rule 1. If the 𝑖th-plus pile has at least one grain and if 𝑀 has a
plus cliff at 𝑖 + 1, then one grain must be added on the 𝑖th-plus
pile:
(5)
βˆ™
:
:
Rule 2. If there is not a plus singleton pile and there is a minus
singleton pile, then the latter must be shifted to the side of the
lowest non empty plus pile:
(6)
: βˆ™
βˆ™ :
(7)
:
Remark 2. (i) Under the hypothesis in Rule 3, the 𝑗th plus
must have at least 2 grains.
(ii) In Rule 2, the lowest non empty plus pile can also be
the invisible column in the place π‘Ÿ + 1. In this case, all the plus
piles are empty and a possible minus singleton pile must be
shifted in the place π‘Ÿ.
4. Main Result
In the sequel, we write 𝑀 → π‘˜ 𝑀󸀠 (or 𝑀󸀠 = 𝑀 → π‘˜ ) to denote
that 𝑀󸀠 is an 𝑛-tuplet of integers obtained from 𝑀 applying the
Rule π‘˜, for π‘˜ = 1, 2, 3, and that ¬(𝑀 → π‘˜ 𝑀󸀠 ) if the condition
𝑀 → π‘˜ 𝑀󸀠 is false. We also set
∇ (𝑀) = {𝑀󸀠 : 𝑀󳨀→π‘˜ 𝑀󸀠 , π‘˜ = 1, 2, 3} .
(8)
Proposition 3. If 𝑀 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ), then ∇(𝑀) ⊆ {𝑀󸀠 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ) :
𝑀󸀠 β‹— 𝑀}.
Proof. Let 𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 |𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ ∈ 𝑆(𝑛, 𝑑, π‘Ÿ), π‘Žπ‘Ÿ+1 = π‘Ÿ+1 (the
invisible pile in the place π‘Ÿ+1), and 𝐷 = 𝐷(𝑀). We distinguish
the three possible cases related to the previous rules.
Case 1. Let us assume that π‘Ÿ ≥ 𝑖 ≥ 1, π‘Žπ‘– =ΜΈ 0, and that 𝑀 has
a plus cliff at 𝑖 + 1. If 𝑀󸀠 = 𝑀 → 1 , then 𝑀󸀠 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Žπ‘–+1 (π‘Žπ‘– +
1)π‘Žπ‘–−1 ⋅ ⋅ ⋅ π‘Ž1 |𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ . It is clear that ‖𝑀󸀠 β€– = 𝑑 because π‘Žπ‘– =ΜΈ 0.
Since there is a plus cliff at 𝑖 + 1, we have π‘Žπ‘–+1 − π‘Žπ‘– ≥ 2; hence,
π‘Žπ‘–+1 ≥ π‘Žπ‘– + 2 > π‘Žπ‘– + 1 > π‘Žπ‘– > π‘Žπ‘–−1 , and this implies that 𝑀󸀠 ∈
𝑆(𝑛, 𝑑, π‘Ÿ). We must show now that 𝑀󸀠 covers 𝑀 in 𝑆(𝑛, 𝑑, π‘Ÿ).
Since 𝑀 and 𝑀󸀠 differ between them only in the place 𝑖 for π‘Žπ‘–
and π‘Žπ‘– + 1, respectively, it is clear that there does not exist an
element 𝑧 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ) such that 𝑀 ⊏ 𝑧 ⊏ 𝑀󸀠 . Hence, 𝑀󸀠 β‹— 𝑀.
Case 2. Let us assume that in 𝐷(𝑀) there is not a plus
singleton pile and that there is a minus singleton pile
(we say 𝐢− (𝑗, 𝐷), for some 1 ≤ 𝑗 ≤ 𝑛 − π‘Ÿ). Since
π‘Žπ‘Ÿ+1 = π‘Ÿ + 1, we can assume that π‘Žπ‘–+1 > 0, π‘Žπ‘– =
0, for some 1 ≤ 𝑖 ≤ π‘Ÿ. This means that 𝑀 has the
following form: 𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Žπ‘–+1 00 ⋅ ⋅ ⋅ 0|0 ⋅ ⋅ ⋅ 0(−1)𝑏𝑗+1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ ,
where π‘Žπ‘–+1 > 1 (otherwise 𝐷(𝑀) has a plus singleton
pile). Applying Rule 2 to 𝑀, we obtain 𝑀󸀠 = 𝑀 → 2 , where
𝑀󸀠 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Žπ‘–+1 10 ⋅ ⋅ ⋅ 0|0 ⋅ ⋅ ⋅ 00𝑏𝑗+1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ . It is clear then that
𝑀󸀠 ∈ 𝑆(𝑛, π‘Ÿ) and ‖𝑀󸀠 β€– = 𝑑 since 𝑀󸀠 is obtained from 𝑀 with
only a shift of the pile-1 to the left in the place 𝑖. Let us note that
the only elements 𝑧1 , 𝑧2 ∈ 𝑆(𝑛, π‘Ÿ) such that 𝑀 ⊏ 𝑧1 ⊏ 𝑀󸀠 and
𝑀 ⊏ 𝑧2 ⊏ 𝑀󸀠 are 𝑧1 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Žπ‘–+1 10 ⋅ ⋅ ⋅ 0|0 ⋅ ⋅ ⋅ 0(−1)𝑏𝑗+1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ
and 𝑧2 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Žπ‘–+1 00 ⋅ ⋅ ⋅ 0|0 ⋅ ⋅ ⋅ 00𝑏𝑗+1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ , but ‖𝑧1 β€– = 𝑑+1
and ‖𝑧2 β€– = 𝑑 − 1; hence, 𝑧1 , 𝑧2 are not elements of 𝑆(𝑛, 𝑑, π‘Ÿ).
This implies that 𝑀󸀠 covers 𝑀 in 𝑆(𝑛, 𝑑, π‘Ÿ).
Case 3. If 1 < 𝑗 ≤ 𝑛 − π‘Ÿ and 𝑀 has a minus cliff at 𝑗, we apply
Rule 3 to 𝑀 on the pile 𝐢− (𝑗, 𝐷) and we obtain 𝑀󸀠 = 𝑀 → 3 ,
4
Journal of Applied Mathematics
where 𝑀󸀠 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 |𝑏1 ⋅ ⋅ ⋅ 𝑏𝑗−1 (𝑏𝑗 +1)𝑏𝑗+1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ . Since 𝑀 has a
minus cliff at 𝑗, we have −𝑏𝑗 +𝑏𝑗−1 = |𝑏𝑗 |−|𝑏𝑗−1 | ≥ 2; therefore,
𝑀󸀠 ∈ 𝑆(𝑛, π‘Ÿ) because 0 ≥ 𝑏𝑗−1 ≥ 𝑏𝑗 + 2 > 𝑏𝑗 + 1 > 𝑏𝑗 > 𝑏𝑗+1 and
‖𝑀󸀠 β€– = 𝑑 since 𝑏𝑗 ≤ −2 implies 𝑏𝑗 + 1 < 0. As in Case 1, we
note that 𝑀󸀠 covers 𝑀 in 𝑆(𝑛, 𝑑, π‘Ÿ) because they differ between
them only for a grain in the place 𝑗.
The main result of this paper is to prove that if 𝑀, 𝑀󸀠 ∈
𝑆(𝑛, 𝑑, π‘Ÿ), then 𝑀󸀠 covers 𝑀 in 𝑆(𝑛, 𝑑, π‘Ÿ) if and only if
𝑀 → π‘˜ 𝑀󸀠 for some π‘˜ ∈ {1, 2, 3}. Obviously this result can be
reformulated as follows.
(∃𝑖 ∈ {1, . . . , π‘Ÿ} such that π‘Žπ‘–+1 > 0, π‘Žπ‘– = 0) AND
(Λ+ (𝑀, 𝑀󸀠 ) \ {𝑖} = Λ− (𝑀, 𝑀󸀠 ) \ {𝑗} = 0, π‘Žπ‘–σΈ€  = 1, 𝑏𝑗󸀠 = 0);
(3) 𝑀 → 3 𝑀󸀠 is equivalent to
Ω3 (𝑀) =ΜΈ 0 AND (∃𝑗 ∈ Ω3 (𝑀) : (𝑏𝑗󸀠 = 𝑏𝑗 + 1) AND
(Λ+ (𝑀, 𝑀󸀠 ) = Λ− (𝑀, 𝑀󸀠 ) \ {𝑗} = 0)).
Negation of Rules 1, 2, and 3.
(1) ¬(𝑀 → 1 𝑀󸀠 ) ⇔ [(1AσΈ€  ) OR (1B)],
where
(1B) is (Ω1 (𝑀) = 0),
(1AσΈ€  ) is [Ω1 (𝑀) =ΜΈ 0 AND (for all 𝑖 ∈ Ω1 (𝑀) (π‘Žπ‘–σΈ€  =ΜΈ π‘Žπ‘– +
1) OR (Λ+ (𝑀, 𝑀󸀠 ) \ {𝑖} =ΜΈ 0 OR Λ− (𝑀, 𝑀󸀠 ) =ΜΈ 0))].
Theorem 4. If 𝑀 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ), then ∇(𝑀) = {𝑀󸀠 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ) :
𝑀󸀠 β‹— 𝑀}.
By Proposition 3, to prove Theorem 4, it suffices to show
that {𝑀󸀠 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ) : 𝑀󸀠 β‹— 𝑀} ⊆ ∇(𝑀). To show this inclusion, we proceed by contradiction as follows. We start with
the following hypotheses:
𝑀, 𝑀󸀠 ∈ 𝑆 (𝑛, 𝑑, π‘Ÿ) ,
𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 | 𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ ,
σΈ€ 
,
𝑀󸀠 = π‘Žπ‘ŸσΈ€  ⋅ ⋅ ⋅ π‘Ž1σΈ€  | 𝑏1σΈ€  ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ
𝑀󸀠 β‹— 𝑀
(2) ¬(𝑀 → 2 𝑀󸀠 ) ⇔ [(2AσΈ€  ) OR (2B) OR (2C) OR (2D)],
where
(2AσΈ€  ) is (for all π‘˜ ∈ {1, . . . , π‘Ÿ} π‘Žπ‘˜ =ΜΈ 1) AND
(∃𝑗 ∈ {1, . . . , 𝑛 − π‘Ÿ} such that 𝑏𝑗 = −1) AND
(∃𝑖 ∈ {1, . . . , π‘Ÿ} such that π‘Žπ‘–+1 > 0, π‘Žπ‘– = 0) AND
((Λ+ (𝑀, 𝑀󸀠 ) \ {𝑖} =ΜΈ 0 OR Λ− (𝑀, 𝑀󸀠 ) \ {𝑗} =ΜΈ 0) OR
π‘Žπ‘–σΈ€  =ΜΈ 1 OR 𝑏𝑗󸀠 =ΜΈ 0);
(2B) is (∃π‘˜ ∈ {1, . . . , π‘Ÿ} such that π‘Žπ‘˜ = 1);
(2C) is (for all 𝑗 ∈ {1, . . . , 𝑛 − π‘Ÿ} 𝑏𝑗 =ΜΈ − 1);
(2D) is (for all 𝑖 ∈ {1, . . . , π‘Ÿ} π‘Žπ‘–+1 = 0 OR π‘Žπ‘– > 0).
(9)
and we also assume that
¬ (𝑀󳨀→π‘˜ 𝑀󸀠 )
for each π‘˜ = 1, 2, 3.
(10)
Then the proof of Theorem 4 will consist into show that
the hypotheses (9) and the conditions (10) always lead to
contradictory cases, so that it must be necessarily verified the
inclusion {𝑀󸀠 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ) : 𝑀󸀠 β‹— 𝑀} ⊆ ∇(𝑀).
We will articulate the proof of Theorem 4 in several cases.
Each possible case will be placed in the form of a proposition
and we will obtain a contradiction in all cases, showing,
hence, the thesis of the theorem. Before proceeding, let us
note that in some cases we also use the alternative notations
𝑀 = 𝑙1 ⋅ ⋅ ⋅ 𝑙𝑛 and 𝑀󸀠 = 𝑙1σΈ€  ⋅ ⋅ ⋅ 𝑙𝑛󸀠 .
We set now
Ω1 (𝑀) = {𝑖 ∈ {1, . . . , π‘Ÿ} : π‘Žπ‘– > 0, π‘Žπ‘–+1 − π‘Žπ‘– ≥ 2} ,
Formalization of Rules 1, 2, and 3.
The description of Rules 1, 2, and 3 in terms of parts of 𝑀
and 𝑀󸀠 is the following:
(1) 𝑀 → 1 𝑀󸀠 is equivalent to
Ω1 (𝑀) =ΜΈ 0 AND (∃𝑖 ∈ Ω1 (𝑀) : (π‘Žπ‘–σΈ€  = π‘Žπ‘– + 1) AND
(Λ+ (𝑀, 𝑀󸀠 ) \ {𝑖} = Λ− (𝑀, 𝑀󸀠 ) = 0));
(for all π‘˜ ∈ {1, . . . , π‘Ÿ} π‘Žπ‘˜ =ΜΈ 1) AND
(∃𝑗 ∈ {1, . . . , 𝑛 − π‘Ÿ} such that 𝑏𝑗 = −1) AND
we set
(1.1b) : 𝑀 = 00 ⋅ ⋅ ⋅ 0 | 𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ ,
(1.2b) : 𝑀 = π‘Ÿ (π‘Ÿ − 1) ⋅ ⋅ ⋅ (π‘˜ + 1) π‘˜ ⋅ ⋅ ⋅ 0 | 𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ
(12)
(13)
Proposition 5. Consider (1B) ⇒ ((1.1b) OR (1.2b)).
(11)
Λ− (𝑀, 𝑀󸀠 ) = {𝑙 ∈ {1, . . . , 𝑛 − π‘Ÿ} : 𝑏𝑙󸀠 > 𝑏𝑙 } .
(2) 𝑀 → 2 𝑀󸀠 is equivalent to
(3B) is (Ω3 (𝑀) = 0);
(3AσΈ€  ) is [Ω3 (𝑀) =ΜΈ 0 AND (for all 𝑗 ∈ Ω3 (𝑀) (𝑏𝑗󸀠 =ΜΈ 𝑏𝑗 +
1) OR (Λ+ (𝑀, 𝑀󸀠 ) =ΜΈ 0 OR Λ− (𝑀, 𝑀󸀠 ) \ {𝑗} =ΜΈ 0))].
for some π‘˜ ∈ {1, . . . , π‘Ÿ} .
Ω3 (𝑀) = {𝑗 ∈ {2, . . . , 𝑛 − π‘Ÿ} : 𝑏𝑗−1 − 𝑏𝑗 ≥ 2} ,
Λ+ (𝑀, 𝑀󸀠 ) = {β„Ž ∈ {1, . . . , π‘Ÿ} : π‘Žβ„ŽσΈ€  > π‘Žβ„Ž } ,
(3) ¬(𝑀 → 3 𝑀󸀠 ) ⇔ [(3AσΈ€  ) OR (3B)],
where
Proof. The condition Ω1 (𝑀) = 0 means that for all 𝑖 ∈
{1, . . . , π‘Ÿ}, π‘Žπ‘– = 0 or 0 ≤ π‘Žπ‘–+1 − π‘Žπ‘– ≤ 1 (i.e., π‘Žπ‘–+1 − π‘Žπ‘– = 0 or
π‘Žπ‘–+1 − π‘Žπ‘– = 1). We have then the following possibilities:
(j) π‘Žπ‘– = 0;
(jj) π‘Žπ‘–+1 − π‘Žπ‘– = 0 (and therefore π‘Žπ‘–+1 = π‘Žπ‘– = 0 since 𝑀 ∈
𝑆(𝑛, π‘Ÿ));
(jjj) π‘Žπ‘–+1 = π‘Žπ‘– + 1.
If π‘Žπ‘– = 0 for all 𝑖 ∈ {1, . . . , π‘Ÿ} then (12) holds. Let us assume
therefore that π‘˜ ∈ {1, . . . , π‘Ÿ} is the smallest index such that
π‘Žπ‘˜ > 0; hence, π‘Žπ‘Ÿ > π‘Žπ‘Ÿ−1 > ⋅ ⋅ ⋅ > π‘Žπ‘˜+1 > π‘Žπ‘˜ > 0 and π‘Žπ‘˜−1 =
⋅ ⋅ ⋅ = π‘Ž1 = 0. This means that for all indexes 𝑖 ∈ {π‘Ÿ, π‘Ÿ−1, . . . , π‘˜+
1, π‘˜} holds necessarily (jjj). In particular, for 𝑖 = π‘Ÿ we have
π‘Ÿ + 1 = π‘Žπ‘Ÿ+1 = π‘Žπ‘Ÿ + 1; hence, π‘Žπ‘Ÿ = π‘Ÿ. Now, if 𝑖 = π‘Ÿ − 1,
π‘Ÿ = π‘Žπ‘Ÿ = π‘Žπ‘Ÿ−1 +1, that is, π‘Žπ‘Ÿ−1 = π‘Ÿ−1. Iterating (13) follows.
Journal of Applied Mathematics
5
(iii) (2C) AND (3.2b) ⇒ absurd,
We set
(2.1d) : 𝑀 = π‘Ÿ (π‘Ÿ − 1) ⋅ ⋅ ⋅ 2 1 | 𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ .
(14)
(iv) (1.1b) AND (3.1b) ⇒ absurd,
(v) (1.2b) AND (3.1b) ⇒ absurd.
Proposition 6. Consider (2D) ⇔ (2.1d).
Proof. By (2D), if we take 𝑖 = π‘Ÿ, we have π‘Žπ‘Ÿ+1 = 0 or π‘Žπ‘Ÿ > 0;
therefore, it is necessarily π‘Žπ‘Ÿ > 0 because by hypothesis, π‘Žπ‘Ÿ+1 =
π‘Ÿ + 1 > 0. Thus, still by (2D), for 𝑖 = π‘Ÿ − 1, we have π‘Ž(π‘Ÿ−1)+1 =
π‘Žπ‘Ÿ = 0 or π‘Žπ‘Ÿ−1 > 0; therefore, it must be π‘Žπ‘Ÿ−1 > 0. By iterating,
we can deduce that π‘Žπ‘Ÿ > 0, π‘Žπ‘Ÿ−1 > 0,. . ., π‘Ž1 > 0; hence, π‘Ž1 = 1,
π‘Ž2 = 2,. . ., π‘Žπ‘Ÿ = π‘Ÿ because 𝑀 is an element of 𝑆(𝑛, π‘Ÿ), that is,
exactly (14). On the other side, if 𝑀 = π‘Ÿ(π‘Ÿ−1) ⋅ ⋅ ⋅ 2 1|𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ ,
then it is immediate to note that (2D) is verified.
Proof. (i), (ii), and (iii) are obvious.
(iv) By (12) and (15), it follows that 𝑀 = 0 ⋅ ⋅ ⋅ 0|0 ⋅ ⋅ ⋅ 0,
absurd.
(v) By Proposition 8, we have 𝑀−σΈ€  = 𝑀− . Then, since
𝑀 ⊏ 𝑀󸀠 , ‖𝑀󸀠 β€– = ‖𝑀‖ and 𝑀−σΈ€  = 𝑀− , by (13) we deduce
that 𝑀+σΈ€  = 𝑀+ . Hence, the absurd 𝑀󸀠 = 𝑀.
Proposition 10. Consider (1.2b) AND (2B) ⇒ (2.1d).
We set
(3.1b) : 𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 | 00 ⋅ ⋅ ⋅ 0,
(3.2b) : 𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 | 00 ⋅ ⋅ ⋅ 0 (−1) (−2) ⋅ ⋅ ⋅ (−𝑙)
for some 𝑙 ∈ {1, . . . , 𝑛 − π‘Ÿ} .
(15)
(16)
Proposition 7. Consider (3B) ⇒ ((3.1b) OR (3.2b)).
Proof. The condition Ω3 (𝑀) = 0 means that for all 𝑗 ∈
{2, . . . , 𝑛−π‘Ÿ}, 0 ≤ 𝑏𝑗−1 −𝑏𝑗 ≤ 1 (i.e., 𝑏𝑗−1 −𝑏𝑗 = 0 or 𝑏𝑗−1 −𝑏𝑖 = 1).
We have then the following possibilities:
(i) 𝑏𝑗−1 − 𝑏𝑗 = 0 (and therefore 𝑏𝑗−1 = 𝑏𝑖 = 0 since 𝑀 ∈
𝑆(𝑛, π‘Ÿ));
(ii) 𝑏𝑗−1 = 𝑏𝑗 + 1.
If (i) holds for all 𝑗 ∈ {2, . . . , 𝑛 − π‘Ÿ}, then (15) holds. We can
assume therefore that π‘˜ ∈ {2, . . . , 𝑛 − π‘Ÿ} is the smallest index
such that (ii) holds. Let us observe then that if in a place 𝑗 ∈
{2, . . . , 𝑛 − π‘Ÿ}, (ii) holds; then also in the places 𝑗 + 1, . . . , 𝑛 − π‘Ÿ,
(ii) holds. In fact, if 𝑏𝑗 = 𝑏𝑗−1 − 1 and by absurd 𝑏𝑗 = 𝑏𝑗+1 =
0, then 𝑏𝑗−1 = 1 which is a contradiction because 0 ≥ 𝑏𝑗−1 .
Therefore, if (ii) holds in 𝑗, then it also holds in 𝑗 + 1, and
therefore, by iteration, it holds for 𝑗+1, 𝑗+2, . . . , 𝑛−π‘Ÿ. Now we
assume at first that π‘˜ = 2. In this case, (ii) holds for all indexes
in {2, . . . , 𝑛 − π‘Ÿ}; therefore, 𝑏1 = 0 or 𝑏1 = −1 and 𝑏2 = 𝑏1 − 1,
𝑏3 = 𝑏2 −1 ⋅ ⋅ ⋅. If 𝑏1 = 0, then 𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 |0(−1)(−2) ⋅ ⋅ ⋅ (−(𝑛−
π‘Ÿ)+1) and if 𝑏1 = −1, then 𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 |(−1)(−2) ⋅ ⋅ ⋅ (−(𝑛−π‘Ÿ));
thus in both cases (ii) holds. Let now π‘˜ ≥ 3. Then π‘π‘˜−1 = π‘π‘˜−2 =
0 (because in the place π‘˜ − 1 (i) holds) and hence 0 = π‘π‘˜−1 =
π‘π‘˜−2 = ⋅ ⋅ ⋅ = 𝑏2 = 𝑏1 . Moreover, since (i) holds in the places
π‘˜, π‘˜+1, . . . , 𝑛−π‘Ÿ, we have π‘π‘˜ = π‘π‘˜−1 −1 = 0−1 = −1; therefore,
π‘π‘˜+1 = π‘π‘˜ − 1 = −2, . . ., 𝑏𝑛−π‘Ÿ = 𝑏𝑛−π‘Ÿ−1 − 1 = −(𝑛 − π‘Ÿ − π‘˜) − 1 = −𝑙,
where 𝑙 = 𝑛 − π‘Ÿ − π‘˜ + 1. This completes the proof.
Proposition 8. Consider (3.1b) ⇒ (𝑀− = 𝑀−σΈ€  =| 00 ⋅ ⋅ ⋅ 0).
Proof. Since 𝑀 ⊏ 𝑀󸀠 , it must be 0 ≥ 𝑏𝑗󸀠 ≥ 𝑏𝑗 , where 𝑏𝑗 = 0 for
all 𝑗 ∈ {1, . . . , 𝑛 − π‘Ÿ} because (15) implies 𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 |00 ⋅ ⋅ ⋅ 0.
Therefore, 𝑏𝑗 = 𝑏𝑗󸀠 = 0 for all 𝑗 ∈ {1, . . . , 𝑛 − π‘Ÿ}; that is, 𝑀󸀠 =
π‘Žπ‘ŸσΈ€  ⋅ ⋅ ⋅ π‘Ž1σΈ€  | 00 ⋅ ⋅ ⋅ 0.
Proposition 9. Consider
(i) (1.1b) AND (2B) ⇒ absurd,
(ii) (1.1b) AND (2.1d) ⇒ absurd,
Proof. By (2B), it follows that in (13) it must be π‘˜ = 1;
therefore, 𝑀 = π‘Ÿ(π‘Ÿ − 1) ⋅ ⋅ ⋅ 2 1|𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ .
Proposition 11. Consider (2.1d) AND (3.2b) ⇒ absurd.
Proof. By (14), it is immediate to observe that 𝑀+σΈ€  = 𝑀+
because 𝑀 ⊏ 𝑀󸀠 . Then by (16), since 𝑀+σΈ€  = 𝑀+ , ‖𝑀−σΈ€  β€– = ‖𝑀− β€–
and 𝑀 ⊏ 𝑀󸀠 , it must also be 𝑀−σΈ€  = |00 ⋅ ⋅ ⋅ 0(−1)(−2) ⋅ ⋅ ⋅ (−𝑙).
Hence, we obtain 𝑀 = 𝑀󸀠 , absurd.
Proposition 12. Consider (2AσΈ€  ) ⇒ absurd.
Proof. We set
(F) [(for all π‘˜ ∈ {1, . . . , π‘Ÿ} π‘Žπ‘˜ =ΜΈ 1) AND (∃𝑗 ∈ {1, . . . , 𝑛 −
π‘Ÿ} such that 𝑏𝑗 = −1) AND (∃𝑖 ∈ {1, . . . , π‘Ÿ} such that
π‘Žπ‘–+1 > 0, π‘Žπ‘– = 0)];
(G) (Λ+ (𝑀, 𝑀󸀠 ) \ {𝑖} =ΜΈ 0 OR Λ− (𝑀, 𝑀󸀠 ) \ {𝑗} =ΜΈ 0);
(H) (π‘Žπ‘–σΈ€  =ΜΈ 1 OR 𝑏𝑗󸀠 =ΜΈ 0).
Then (2AσΈ€  ) is equivalent to ((F) AND (G)) OR ((F) AND (H)).
We show at first that ((F) AND (H)) leads to a contradiction.
Next we show that ((F) AND (G) AND ¬(H)) also leads to an
absurd. This will prove the thesis;
((F) AND (H)) ⇒ absurd.
By (F), it follows that 𝑀 satisfies the hypothesis of Rule 2;
therefore, by Proposition 3 we can take an element 𝑧 ∈
𝑆(𝑛, 𝑑, π‘Ÿ) such that 𝑧 = 𝑀 → 2 . As in the proof of the Case 2 in
Proposition 3, it results that 𝑀 and 𝑧 must have the following
form:
𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Žπ‘–+1 00 ⋅ ⋅ ⋅ 0 | 0 ⋅ ⋅ ⋅ 0 (−1) 𝑏𝑗+1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ ,
(17)
𝑧 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Žπ‘–+1 10 ⋅ ⋅ ⋅ 0 | 0 ⋅ ⋅ ⋅ 00𝑏𝑗+1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ ,
(18)
where π‘Žπ‘–+1 > 1 and 𝑏𝑗+1 < −1. We distinguish now several
cases.
(I) 𝑏𝑗󸀠 =ΜΈ 0. Then 𝑏𝑗󸀠 ≤ −1, and also 𝑏𝑗󸀠 ≥ 𝑏𝑗 = −1 because
𝑀 ⊏ 𝑀󸀠 ; hence 𝑏𝑗󸀠 = −1.
We note now that
σΈ€ 
π‘Žπ‘ŸσΈ€  > 0, . . . , π‘Žπ‘–+1
>0
(19)
6
Journal of Applied Mathematics
σΈ€ 
because π‘Žπ‘ŸσΈ€  ≥ π‘Žπ‘Ÿ > 0, . . . , π‘Žπ‘–+1
≥ π‘Žπ‘–+1 > 0. Moreover, since
σΈ€ 
𝑏𝑗 = −1, it follows that
σΈ€ 
σΈ€ 
−1 = 𝑏𝑗󸀠 > 𝑏𝑗+1
> ⋅ ⋅ ⋅ > 𝑏𝑛−π‘Ÿ
(20)
because 𝑀󸀠 ∈ 𝑆(𝑛, π‘Ÿ). Now, since 𝑀 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ), from the form
of 𝑀 it results that 𝑑 = (π‘Ÿ − 𝑖) + (𝑛 − π‘Ÿ − 𝑗 + 1). Moreover, since
also 𝑀󸀠 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ), we have ‖𝑀󸀠 β€– = 𝑑 = (π‘Ÿ − 𝑖) + (𝑛 − π‘Ÿ − 𝑗 + 1).
Therefore, since in (19), there are exactly π‘Ÿ − 𝑖 inequalities and
in (20) there are exactly 𝑛 − π‘Ÿ − 𝑗 + 1 inequalities; it follows
that
σΈ€ 
σΈ€ 
σΈ€ 
00 ⋅ ⋅ ⋅ 0 | 0 ⋅ ⋅ ⋅ 0 (−1) 𝑏𝑗+1
⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ
.
𝑀󸀠 = π‘Žπ‘ŸσΈ€  ⋅ ⋅ ⋅ π‘Žπ‘–+1
(21)
(II) π‘Žπ‘–σΈ€  = 0. Since 𝑀 ⊏ 𝑀󸀠 , by (17) we have two possible
cases: 𝑀󸀠 has the form (21) or
σΈ€ 
σΈ€ 
σΈ€ 
𝑀󸀠 = π‘Žπ‘ŸσΈ€  ⋅ ⋅ ⋅ π‘Žπ‘–+1
00 ⋅ ⋅ ⋅ 0 | 0 ⋅ ⋅ ⋅ 00 𝑏𝑗+1
⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ
(22)
σΈ€ 
σΈ€ 
≥ π‘Žπ‘–+1 > 1 and 0 ≥ 𝑏𝑗+1
. If (22) holds, then ‖𝑀󸀠 β€– ≤
with π‘Žπ‘–+1
(π‘Ÿ − 𝑖) + (𝑛 − π‘Ÿ − 𝑗) = 𝑑 − 1, which is absurd because 𝑀󸀠 ∈
𝑆(𝑛, 𝑑, π‘Ÿ); hence it must hold necessarily (21).
(III) π‘Žπ‘–σΈ€  > 1. Since 𝑀 ⊏ 𝑀󸀠 and 𝑑 = ‖𝑀‖ = ‖𝑀󸀠 β€–, by (17) it
follows that
σΈ€ 
σΈ€ 
σΈ€ 
𝑀󸀠 = π‘Žπ‘ŸσΈ€  ⋅ ⋅ ⋅ π‘Žπ‘–+1
π‘Žπ‘–σΈ€  0 ⋅ ⋅ ⋅ 0 | 0 ⋅ ⋅ ⋅ 00 𝑏𝑗+1
⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ
.
(23)
Then from the hypothesis π‘Žπ‘–σΈ€  ≥ 2 and by (18), we have 𝑀 ⊏
𝑧 ⊏ 𝑀󸀠 , which is a contradiction.
Hence the unique possibility is that 𝑀󸀠 has the form (21).
Since 𝑀󸀠 β‹— 𝑀, we have π‘Žπ‘˜σΈ€  ≥ π‘Žπ‘˜ for π‘˜ = π‘Ÿ, . . . , 𝑖 + 2, 𝑖 + 1 and
𝑏𝑙󸀠 ≥ 𝑏𝑙 for 𝑙 = 𝑗 + 1, 𝑗 + 2, . . . , 𝑛 − π‘Ÿ. Moreover, at least one
inequality of the type π‘Žπ‘˜σΈ€  > π‘Žπ‘˜ or the type 𝑏𝑙󸀠 > 𝑏𝑙 must hold.
We distinguish now the several cases.
(Δ1) We assume that a unique inequality of the type π‘Žπ‘σΈ€  >
π‘Žπ‘ holds for some 𝑝 = π‘Ÿ, . . . , 𝑖 + 2, 𝑖 + 1. Therefore, we have
π‘Žπ‘˜σΈ€  = π‘Žπ‘˜ for all π‘˜ =ΜΈ 𝑝 and 𝑏𝑙󸀠 = 𝑏𝑙 for all 𝑙 = 𝑗 + 1, 𝑗 + 2, . . . , 𝑛 − π‘Ÿ.
σΈ€ 
Then, since π‘Žπ‘+1 = π‘Žπ‘+1
> π‘Žπ‘σΈ€  > π‘Žπ‘ > 0 (let us note that the
previous inequalities also hold if 𝑝 = π‘Ÿ because π‘Žπ‘Ÿ+1 = π‘Ÿ + 1 =
σΈ€ 
), we have that π‘Žπ‘+1 − π‘Žπ‘ ≥ 2 and π‘Žπ‘ > 0. Hence 𝑀 satisfies
π‘Žπ‘Ÿ+1
the hypotheses of Rule 1 in the place 𝑝 and so 𝑀 → 1 𝑒, where
𝑒 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Žπ‘+1 (π‘Žπ‘ + 1) π‘Žπ‘−1 ⋅ ⋅ ⋅ π‘Žπ‘–+1 0 ⋅ ⋅ ⋅ 0 |
V = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Žπ‘+1 (π‘Žπ‘ + 1) π‘Žπ‘−1 ⋅ ⋅ ⋅ π‘Žπ‘–+1 0 ⋅ ⋅ ⋅ 0 |
0 ⋅ ⋅ ⋅ 0 (−1) 𝑏𝑗+1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ .
(25)
Since π‘Žπ‘σΈ€  > π‘Žπ‘ is the first strict inequality by left, we have
σΈ€ 
σΈ€ 
= π‘Žπ‘+1 (also if 𝑝 = π‘Ÿ); hence π‘Žπ‘+1 = π‘Žπ‘+1
> π‘Žπ‘σΈ€  ≥
π‘Žπ‘+1
π‘Žπ‘ +1 > π‘Žπ‘ > π‘Žπ‘−1 because π‘Žπ‘ > 0; this implies that V ∈ 𝑆(𝑛, π‘Ÿ);
therefore V ∈ 𝑆(𝑛, 𝑑, π‘Ÿ) because β€–Vβ€– = ‖𝑀‖ = 𝑑. Moreover, it is
clear that 𝑀 ⊏ V. Let us observe now that V ⊏ 𝑀󸀠 ; in fact, by
hypothesis, 𝑀󸀠 must have at least another place π‘ž =ΜΈ 𝑝 where
π‘Žπ‘žσΈ€  > π‘Žπ‘ž or π‘π‘žσΈ€  > π‘π‘ž and V differs by 𝑀 only in the place 𝑝. This
shows that 𝑀 ⊏ V ⊏ 𝑀󸀠 , with V ∈ 𝑆(𝑛, 𝑑, π‘Ÿ), which is absurd
because 𝑀󸀠 β‹— 𝑀.
(Δ4) If at least two strict inequalities among the π‘Žπ‘˜σΈ€  ≥ π‘Žπ‘˜
and 𝑏𝑙󸀠 ≥ 𝑏𝑙 hold and the first (by left) of such strict inequalities
is 𝑏𝑝󸀠 > 𝑏𝑝 for some 𝑝 ∈ {𝑗 + 1, 𝑗 + 2, . . . , 𝑛 − π‘Ÿ}, then we deduce
a contradiction in the same way as in the previous case.
This concludes the proof that ((F) AND (H)) ⇒ absurd.
To complete the proof we show now that ((F) AND (G) AND
¬(H)) ⇒ absurd.
Since (H) is false, we have π‘Žπ‘–σΈ€  = 1 and 𝑏𝑗󸀠 = 0; thereσΈ€ 
fore, 𝑀 has the form (17) and 𝑀󸀠 = π‘Žπ‘ŸσΈ€  ⋅ ⋅ ⋅ π‘Žπ‘–+1
10 ⋅ ⋅ ⋅ 0|0 ⋅ ⋅ ⋅
σΈ€ 
σΈ€ 
⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ
. We take 𝑧 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ) as in (18), so that 𝑀
00𝑏𝑗+1
and 𝑧 differ between them only in the places 𝑖,𝑗 and 𝑀 ⊏ 𝑧.
σΈ€ 
σΈ€ 
σΈ€ 
10 ⋅ ⋅ ⋅ 0|0 ⋅ ⋅ ⋅ 00𝑏𝑗+1
⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ
,
Moreover, since 𝑀󸀠 = π‘Žπ‘ŸσΈ€  ⋅ ⋅ ⋅ π‘Žπ‘–+1
σΈ€ 
by (G), (18), and (17) it follows that 𝑧 ⊏ 𝑀 . Therefore, we
obtain 𝑀 ⊏ 𝑧 ⊏ 𝑀󸀠 that is in contrast with the hypothesis
𝑀󸀠 β‹— 𝑀.
We set now
(1A) : Ω1 (𝑀) =ΜΈ 0,
∀𝑖 ∈ Ω1 (𝑀) π‘Žπ‘–σΈ€  =ΜΈ π‘Žπ‘– + 1,
(26)
(3A) : Ω3 (𝑀) =ΜΈ 0,
∀𝑗 ∈ Ω1 (𝑀) 𝑏𝑗󸀠 =ΜΈ 𝑏𝑗 + 1.
(27)
Proposition 13. Consider
0 ⋅ ⋅ ⋅ 0 (−1) 𝑏𝑗+1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ
σΈ€ 
σΈ€ 
σΈ€ 
(= π‘Žπ‘ŸσΈ€  ⋅ ⋅ ⋅ π‘Žπ‘+1
(π‘Žπ‘ + 1) π‘Žπ‘−1
⋅ ⋅ ⋅ π‘Žπ‘–+1
0⋅⋅⋅0 |
as in the proof of the previous case by using Rule 3 instead of
Rule 1.
(Δ3) We assume now that at least two strict inequalities
among the π‘Žπ‘˜σΈ€  ≥ π‘Žπ‘˜ and 𝑏𝑙󸀠 ≥ 𝑏𝑙 hold, and let us also assume
that the first (by left) of such strict inequalities is π‘Žπ‘σΈ€  > π‘Žπ‘ for
some 𝑝 ∈ {π‘Ÿ, . . . , 𝑖 + 1}. We consider then the following (𝑛, π‘Ÿ)string:
(24)
σΈ€ 
σΈ€ 
⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ
),
0 ⋅ ⋅ ⋅ 0 (−1) 𝑏𝑗+1
with 𝑒 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ). Now, if π‘Žπ‘σΈ€  = π‘Žπ‘ + 1, it results that
𝑀󸀠 = 𝑒; therefore, 𝑀 → 1 𝑀󸀠 , which is a contradiction with
the hypothesis ¬(𝑀 → 1 𝑀󸀠 ). Otherwise, if π‘Žπ‘ + 1 < π‘Žπ‘σΈ€  , then
we have 𝑀 ⊏ 𝑒 ⊏ 𝑀󸀠 , which is absurd because 𝑀󸀠 β‹— 𝑀.
(Δ2) If a unique inequality of the type π‘π‘žσΈ€  > π‘π‘ž holds for
some π‘ž = 𝑗 + 1, 𝑗 + 2, . . . , 𝑛 − π‘Ÿ, we deduce a contradiction
(i) (1AσΈ€  ) ⇔ (1A),
(ii) (3AσΈ€  ) ⇔ (3A).
Proof. (i) The implication ⇐ is obvious. We assume now that
(1AσΈ€  ) holds and there exists 𝑖 ∈ Ω1 (𝑀) such that π‘Žπ‘–σΈ€  = π‘Žπ‘– + 1
and also that Λ+ (𝑀, 𝑀󸀠 )\{𝑖} =ΜΈ 0 (i.e., there exists β„Ž =ΜΈ 𝑖 such that
π‘Žβ„ŽσΈ€  > π‘Žβ„Ž ) or Λ− (𝑀, 𝑀󸀠 ) =ΜΈ 0 (i.e., there exists 𝑙 such that 𝑏𝑙󸀠 > 𝑏𝑙 ).
Since 𝑖 ∈ Ω1 (𝑀), we can apply Rule 1 to 𝑀 and we can take
𝑧 = 𝑀 → 1 . By Proposition 3, we know that 𝑧 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ) and
𝑧 β‹— 𝑀. Moreover, since π‘Žβ„ŽσΈ€  > π‘Žβ„Ž (β„Ž =ΜΈ 𝑖) or 𝑏𝑙󸀠 > 𝑏𝑙 , we also
have 𝑧 ⊏ 𝑀󸀠 ; therefore, 𝑀 ⊏ 𝑧 ⊏ 𝑀󸀠 , which is in contrast with
the hypothesis 𝑀󸀠 β‹— 𝑀. Hence this proves that the condition
Journal of Applied Mathematics
7
π‘Žπ‘–σΈ€  =ΜΈ π‘Žπ‘– + 1 must be verified for all 𝑖 ∈ Ω1 (𝑀), which is exactly
(26).
(ii) The proof is similar to (i) applying Rule 3 to 𝑀 instead
of Rule 1.
We set
(1a) : Ω1 (𝑀) =ΜΈ 0,
∀𝑖 ∈ Ω1 (𝑀) π‘Žπ‘–σΈ€  = π‘Žπ‘– ,
(28)
(3a) : Ω3 (𝑀) =ΜΈ 0,
∀𝑗 ∈ Ω3 (𝑀) 𝑏𝑗󸀠 = 𝑏𝑗 .
(29)
Proposition 14. Consider
Lemma 17. Assume that (3a) holds. If π‘ž ∈ {2, 3, . . . , 𝑛 − π‘Ÿ − 1},
σΈ€ 
= π‘π‘ž+1 .
π‘π‘žσΈ€  = π‘π‘ž < 0, then π‘π‘ž+1
(i) (1A) ⇒ (1a),
(ii) (3A) ⇒ (3a).
Proof. (i) By (1A), we have Ω1 (𝑀) =ΜΈ 0, and by the definition of
Ω1 (𝑀) this is equivalent to say that 𝑀 satisfies the hypotheses
of Rule 1 for some 𝑖 ∈ {1, . . . , π‘Ÿ}. Therefore, by Proposition 3,
there exists an element 𝑧𝑖 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ) such that 𝑀 → 1 𝑧𝑖 ,
where 𝑧𝑖 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Žπ‘–+1 (π‘Žπ‘– + 1)π‘Žπ‘–−1 ⋅ ⋅ ⋅ π‘Ž1 |𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ and obviously
𝑀 ⊏ 𝑧𝑖 . Since π‘Žπ‘–σΈ€  =ΜΈ π‘Žπ‘– + 1, we distinguish two cases.
(a) π‘Žπ‘–σΈ€  > π‘Žπ‘– + 1. In this case 𝑀 ⊏ 𝑧𝑖 ⊏ 𝑀󸀠 , that is in contrast
with the hypothesis 𝑀󸀠 β‹— 𝑀.
(b) π‘Žπ‘–σΈ€  < π‘Žπ‘– + 1. In this case π‘Žπ‘–σΈ€  ≤ π‘Žπ‘– ; hence π‘Žπ‘–σΈ€  = π‘Žπ‘– because
π‘Žπ‘–σΈ€  ≥ π‘Žπ‘– from the hypothesis 𝑀󸀠 β‹— 𝑀.
(ii) The proof of this case is similar to (i).
Lemma 15. Assume that (1a) holds. If π‘ž ∈ {2, 3, . . . , π‘Ÿ, π‘Ÿ + 1},
σΈ€ 
= π‘Žπ‘ž−1 .
π‘Žπ‘žσΈ€  = π‘Žπ‘ž , and π‘Žπ‘ž−1 > 0, then π‘Žπ‘ž−1
Proof. If π‘ž−1 ∈ Ω1 (𝑀), then the thesis holds by the definition
of (1a) in (28). We assume now that π‘ž − 1 ∉ Ω1 (𝑀). Then
π‘Žπ‘ž−1 = 0 or 0 ≤ π‘Ž(π‘ž−1)+1 − π‘Žπ‘ž−1 = π‘Žπ‘ž − π‘Žπ‘ž−1 ≤ 1, and since
π‘Žπ‘ž−1 > 0, we necessarily have 0 ≤ π‘Žπ‘ž − π‘Žπ‘ž−1 ≤ 1. Now, if
π‘Žπ‘ž − π‘Žπ‘ž−1 = 0, we have π‘Žπ‘ž = π‘Žπ‘ž−1 = 0 since 𝑀 ∈ 𝑆(𝑛, π‘Ÿ),
which is in contradiction with π‘Žπ‘ž−1 > 0. Hence, it must be
σΈ€ 
π‘Žπ‘ž − π‘Žπ‘ž−1 = 1; that is, π‘Žπ‘ž = π‘Žπ‘ž−1 + 1. Moreover, π‘Žπ‘žσΈ€  > π‘Žπ‘ž−1
σΈ€ 
σΈ€ 
because π‘Žπ‘ž−1 ≥ π‘Žπ‘ž−1 > 0 and 𝑀 ∈ 𝑆(𝑛, π‘Ÿ). Therefore, since
σΈ€ 
, which implies
π‘Žπ‘žσΈ€  = π‘Žπ‘ž , we have π‘Žπ‘ž−1 + 1 = π‘Žπ‘ž = π‘Žπ‘žσΈ€  > π‘Žπ‘ž−1
σΈ€ 
σΈ€ 
π‘Žπ‘ž−1 + 1 > π‘Žπ‘ž−1 ≥ π‘Žπ‘ž−1 because 𝑀 ⊏ 𝑀 . By subtracting π‘Žπ‘ž−1
from each term of the previous inequalities, we obtain 1 >
σΈ€ 
σΈ€ 
− π‘Žπ‘ž−1 ≥ 0; therefore, it must be π‘Žπ‘ž−1
= π‘Žπ‘ž−1 .
π‘Žπ‘ž−1
We set
(1.1a) : ∃𝑙 ∈ {π‘Ÿ, . . . , 2}
such that
π‘Žπ‘™ > 0,
𝑀 = π‘Žπ‘Ÿ π‘Žπ‘Ÿ−1 ⋅ ⋅ ⋅ π‘Žπ‘™+1 π‘Žπ‘™ 0 ⋅ ⋅ ⋅ 0 | 𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ ,
σΈ€ 
𝑀 =
𝑀 ∈ 𝑆(𝑛, π‘Ÿ), it must be π‘Ž1 = 0 or π‘Ž1 = 1. If π‘Ž1 = 1, then π‘Ž2 =
2, . . . , π‘Žπ‘Ÿ = π‘Ÿ, which is in contrast with π‘Žπ‘–+1 − π‘Žπ‘– ≥ 2. Hence
π‘Ž1 = 0 and therefore there exists 𝑙 ∈ 2, . . . , π‘Ÿ such that π‘Žπ‘™ > 0
and π‘Žπ‘™−1 = ⋅ ⋅ ⋅ = π‘Ž1 = 0. We take π‘ž = π‘Ÿ + 1; that is, π‘ž − 1 = π‘Ÿ.
Since π‘Žπ‘žσΈ€  = π‘Žπ‘ž = π‘Ÿ + 1 and π‘Žπ‘ž−1 = π‘Žπ‘Ÿ ≥ π‘Žπ‘™ > 0, by Lemma 15 it
σΈ€ 
= π‘Žπ‘ž−1 ; that is, π‘Žπ‘ŸσΈ€  = π‘Žπ‘Ÿ . Let now π‘ž = π‘Ÿ. If π‘Ÿ = 𝑙,
follows that π‘Žπ‘ž−1
then we have the thesis otherwise π‘ž = π‘Ÿ > 𝑙; that is, π‘ž − 1 ≥ 𝑙;
therefore π‘Žπ‘ž−1 ≥ π‘Žπ‘™ > 0 because 𝑀 ∈ 𝑆(𝑛, π‘Ÿ). Then since π‘Žπ‘žσΈ€  =
σΈ€ 
=
π‘Žπ‘ŸσΈ€  = π‘Žπ‘Ÿ = π‘Žπ‘ž , we can apply Lemma 15 and we deduce π‘Žπ‘ž−1
σΈ€ 
π‘Žπ‘ž−1 ; that is, π‘Žπ‘Ÿ−1 = π‘Žπ‘Ÿ−1 . By iteration the thesis follows.
σΈ€ 
π‘Žπ‘Ÿ π‘Žπ‘Ÿ−1 ⋅ ⋅ ⋅ π‘Žπ‘™+1 π‘Žπ‘™ π‘Žπ‘™−1
⋅ ⋅ ⋅ π‘Ž1σΈ€ 
|
𝑏1σΈ€ 
(30)
σΈ€ 
⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ
.
Proposition 16. Consider (1a) ⇒ (1.1a).
Proof. By (28), there exists at least one element 𝑖 ∈ Ω1 (𝑀). In
this place we have π‘Žπ‘– > 0 and moreover π‘Žπ‘–+1 − π‘Žπ‘– ≥ 2. Since
Proof. If π‘ž + 1 ∈ Ω3 (𝑀), then the thesis holds by definition
of (3a) in (29). We assume now that π‘ž + 1 ∉ Ω3 (𝑀). Now, if
π‘π‘ž − π‘π‘ž+1 = 0, we have π‘π‘ž = π‘π‘ž+1 = 0 since 𝑀 ∈ 𝑆(𝑛, π‘Ÿ), which is
in contradiction with the hypothesis π‘π‘ž < 0. Hence it must be
π‘π‘ž − π‘π‘ž+1 = 1; that is, 0 > π‘π‘ž = π‘π‘ž+1 + 1; therefore 0 > π‘π‘ž+1 + 1 =
σΈ€ 
π‘π‘ž = π‘π‘žσΈ€  > π‘π‘ž+1
≥ π‘π‘ž+1 since 𝑀󸀠 ∈ 𝑆(𝑛, π‘Ÿ) and 𝑀 ⊏ 𝑀󸀠 . The
σΈ€ 
last inequalities provide π‘π‘ž+1 + 1 > π‘π‘ž+1
≥ π‘π‘ž+1 . By subtracting
π‘π‘ž+1 from each term of the previous inequalities, we obtain
σΈ€ 
σΈ€ 
1 > π‘π‘ž+1
− π‘π‘ž+1 ≥ 0; therefore it must be π‘π‘ž+1
= π‘π‘ž+1 .
Proposition 18. If (3a) holds, then 𝑀 has the following form:
𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 |0 ⋅ ⋅ ⋅ 0 𝑏𝑙 𝑏𝑙+1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ , with 𝑏𝑙 < 0 and 𝑙 ≥ 2.
Moreover, if 𝑏𝑙 ≤ −2, we also have 𝑀−σΈ€  = 𝑀− .
Proof. By (3a) in (29), we know that Ω3 (𝑀) =ΜΈ 0. Moreover, if
𝑗 ∈ Ω3 (𝑀), then 𝑗 ≥ 2 and 𝑏𝑗 < 0 because 𝑏𝑗−1 − 𝑏𝑗 ≥ 2.
Therefore, we can take the minimum 𝑙 ∈ 2, . . . , 𝑛 − π‘Ÿ such
that 𝑏𝑙 < 0; hence 𝑏𝑙−1 = ⋅ ⋅ ⋅ = 𝑏1 = 0 and this means that
𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 |0 ⋅ ⋅ ⋅ 0𝑏𝑙 𝑏𝑙+1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ . Assume now that 𝑏𝑙 ≤ −2.
Since 𝑏𝑙−1 = 0, we have 𝑏𝑙−1 − 𝑏𝑙 ≤ 2; therefore 𝑙 ∈ Ω3 (𝑀). By
(3a), follows then that 𝑏𝑙󸀠 = 𝑏𝑙 < 0. Now, if 𝑙 = 𝑛 − π‘Ÿ, the thesis
is true, otherwise, if 𝑙 ∈ {2, . . . , 𝑛 − π‘Ÿ − 1}, by Lemma 17 we
σΈ€ 
= 𝑏𝑙+1 < 0. Again, if 𝑙 + 1 = 𝑛 − π‘Ÿ, the thesis is true;
have 𝑏𝑙+1
otherwise, if 𝑙 + 1 ∈ {2, . . . , 𝑛 − π‘Ÿ − 1}, by Lemma 17 we have
σΈ€ 
= 𝑏𝑙+2 < 0. Proceeding repeatedly in this way we obtain
𝑏𝑙+2
σΈ€ 
σΈ€ 
= 𝑏𝑙+1 , . . . , 𝑏𝑛−π‘Ÿ
= 𝑏𝑛−π‘Ÿ . Finally, since 𝑏1 = ⋅ ⋅ ⋅ =
𝑏𝑙󸀠 = 𝑏𝑙 , 𝑏𝑙+1
σΈ€ 
σΈ€ 
= 0. This
𝑏𝑙−1 = 0 and 𝑀 ⊏ 𝑀 , we also have 𝑏1σΈ€  = ⋅ ⋅ ⋅ = 𝑏𝑙−1
σΈ€ 
proves that 𝑀− = 𝑀− .
Proposition 19. Consider ((2C) AND (3a)) ⇒ 𝑀−σΈ€  = 𝑀− .
Proof. (2C) implies that 𝑀 has the form 𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 |0 ⋅ ⋅ ⋅
0 𝑏𝑙 𝑏𝑙+1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ , with 𝑏𝑙 ≤ −2. Since (3a) holds, by
Proposition 18, the thesis follows.
Proposition 20. Consider ((1.1a) AND (2C) AND (3a)) ⇒
absurd.
Proof. By Proposition 19, we have that 𝑀−σΈ€  = 𝑀− . Moreover,
by (1.1a) in (30), we also have 𝑀+σΈ€  = 𝑀+ because ‖𝑀󸀠 β€– = ‖𝑀‖ =
𝑑 and 𝑀−σΈ€  = 𝑀− ; therefore, 𝑀 = 𝑀󸀠 . This is absurd because
𝑀󸀠 β‹— 𝑀 by hypothesis.
8
Journal of Applied Mathematics
Proposition 21. Consider ((1.1b) AND (2C) AND (3a)) ⇒
absurd.
Proof. By (1.1b) in (12), we know that 𝑀 = 0 ⋅ ⋅ ⋅ 0|𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ ,
and this implies ‖𝑀+ β€– = 0 and ‖𝑀− β€– = 𝑑. Moreover, by
Proposition 19, 𝑀−σΈ€  = 𝑀− ; therefore it follows that ‖𝑀−σΈ€  β€– = 𝑑;
hence ‖𝑀+σΈ€  β€– = 0 since 𝑀󸀠 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ). This shows that 𝑀 = 𝑀󸀠
and this is absurd because 𝑀 β‹— 𝑀󸀠 .
Proposition 22. Consider ((1.1a) AND (2B)) ⇒ 𝑀+σΈ€  = 𝑀+ .
Proof. By (1.1a) in (30) and (2B), it follows that 𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅
σΈ€ 
⋅ ⋅ ⋅ π‘Ž1σΈ€  |𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ .
π‘Žπ‘™+1 10 ⋅ ⋅ ⋅ 0|𝑏1 ⋅ ⋅ ⋅ 𝑏𝑛−π‘Ÿ and 𝑀󸀠 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Žπ‘™+1 1π‘Žπ‘™−1
Now, since 𝑀󸀠 ∈ 𝑆(𝑛, π‘Ÿ) and π‘Žπ‘™ = π‘Žπ‘™σΈ€  = 1, we have that 1 = π‘Žπ‘™σΈ€  >
σΈ€ 
σΈ€ 
π‘Žπ‘™−1
≥ ⋅ ⋅ ⋅ ≥ π‘Ž1σΈ€  ≥ 0; that is, π‘Žπ‘™−1
= ⋅ ⋅ ⋅ = π‘Ž1σΈ€  = 0. Therefore
σΈ€ 
𝑀+ = 𝑀+ .
Proposition 23. Consider ((1.1a) AND (2B) AND (3.1b)) ⇒
absurd.
Proof. By Proposition 22 𝑀+σΈ€  = 𝑀+ and by Proposition 8 𝑀 =
𝑀−σΈ€  ; hence 𝑀 = 𝑀󸀠 and this is a contradiction.
Proposition 24. Consider ((1.1a) AND (3.1b)) ⇒ absurd.
Proof. By Proposition 8, 𝑀 = 𝑀−σΈ€  = |00 ⋅ ⋅ ⋅ 0; hence ‖𝑀+σΈ€  β€– =
‖𝑀+ β€– = 𝑑 because 𝑀, 𝑀󸀠 ∈ 𝑆(𝑛, 𝑑, π‘Ÿ). Moreover, by (1.1a) in
(30), it follows that
𝑀+σΈ€  = 𝑀+ = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Žπ‘™+1 π‘Žπ‘™ 0 ⋅ ⋅ ⋅ 0.
(31)
Proof. By Proposition 22, 𝑀+σΈ€  = 𝑀+ . Then by (3.2b) in (16),
since 𝑀+σΈ€  = 𝑀+ , ‖𝑀−σΈ€  β€– = ‖𝑀− β€–, and 𝑀 ⊏ 𝑀󸀠 , it must also be
𝑀−σΈ€  =| 00 ⋅ ⋅ ⋅ 0 (−1) (−2) ⋅ ⋅ ⋅ (−𝑙) .
Hence we obtain 𝑀 = 𝑀󸀠 , absurd.
Proposition 27. Consider ((1.2b) AND (2C) AND (3a)) ⇒
absurd.
Proof. By Proposition 19, 𝑀−σΈ€  = 𝑀− . Then, since 𝑀 ⊏ 𝑀󸀠 ,
‖𝑀󸀠 β€– = ‖𝑀‖, and 𝑀−σΈ€  = 𝑀− , by (1.2b) in (13) we deduce that
𝑀+σΈ€  = 𝑀+ . Hence the absurd 𝑀󸀠 = 𝑀.
Proposition 28. Consider ((2.1d) AND (3a)) ⇒ absurd.
Proof. Since 𝑀 ⊏ 𝑀󸀠 by (2.1d) in (14), it follows that
𝑀+σΈ€  = 𝑀+ . Then, since (3a) in (29) holds, the same proof of
Proposition 25 leads to the absurd 𝑀 = 𝑀󸀠 .
We complete now the proof of our main result; that is,
we provide three syntactic trees where each possible path
from the roots to the leaves leads to a contradiction. In the
following we draw each tree and next we also write all the
implication chains for each path from the roots to the leafs.
We recall at first some useful implications and equivalences already proved in the previous section:
(Θ1A) (1AσΈ€  ) ⇔ [by Proposition 13 (i)] (1A) ⇒ [by
Proposition 14] (1a) ⇒ [by Proposition 16] (1.1a)
Therefore, 𝑀 = 𝑀󸀠 , which is absurd.
(Θ2A) (2AσΈ€  ) ⇒ [by Proposition 12] absurd
Proposition 25. Consider ((1.1a) AND (2B) AND (3a)) ⇒
absurd.
(Θ2D) (2D) ⇔ [by Proposition 6] (2.1d)
Proof. By Proposition 18, we have 𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 |0 ⋅ ⋅ ⋅ 0𝑏𝑙 𝑏𝑙+1 ⋅ ⋅ ⋅
𝑏𝑛−π‘Ÿ with 𝑏𝑙 < 0 and 𝑙 ≥ 2. By Proposition 22, 𝑀+σΈ€  = 𝑀+ . Now,
if 𝑏𝑙 ≤ −2, by Proposition 18 we have 𝑀−σΈ€  = 𝑀− and therefore
𝑀󸀠 = 𝑀, which is absurd. So we can assume that 𝑏𝑙 = −1.
Since 𝑀+σΈ€  = 𝑀+ , ‖𝑀󸀠 β€– = ‖𝑀‖ = 𝑑 and 𝑀 ⊏ 𝑀󸀠 , it follows that
𝑏𝑙󸀠 cannot be 0; otherwise, it is ‖𝑀󸀠 β€– < 𝑑. This implies that it
must be 𝑏𝑙󸀠 = −1. We distinguish now two cases.
(Θ3A) (3AσΈ€  ) ⇔ [by Proposition 13 (ii)] (3A) ⇒ [by
Proposition 14] (3a) ⇒ [see statement of
Proposition 18] Proposition 18.
The first tree that we consider is that having the condition
(1AσΈ€  ) as root (we denote such tree with TREE(1AσΈ€  )).
TREE (1AσΈ€  ):
1Aσ³°€
𝑏𝑙󸀠
Case 1. If 𝑙 ∈ {2, . . . , 𝑛−π‘Ÿ−1}, since 𝑏𝑙 = = −1, by Lemma 17
σΈ€ 
(with π‘ž = 𝑙) we obtain 𝑏𝑙+1 = 𝑏𝑙+1
< 0. Now, if 𝑙 + 1 = 𝑛 − π‘Ÿ,
σΈ€ 
then 𝑀− = 𝑀− , and as before this provides a contradiction. If
𝑙 + 1 < 𝑛 − π‘Ÿ, we can still apply Lemma 17 (with π‘ž = 𝑙 + 1) and
σΈ€ 
< 0. Proceeding in this way in all
therefore have 𝑏𝑙+2 = 𝑏𝑙+2
cases we always obtain 𝑀− = 𝑀−σΈ€  and hence is absurd.
Case 2. If 𝑙 = 𝑛 − π‘Ÿ, then 𝑀 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 | 0 ⋅ ⋅ ⋅ 0(−1), and since
‖𝑀󸀠 β€– = ‖𝑀‖ = 𝑑, 𝑀 ⊏ 𝑀󸀠 , and 𝑀+σΈ€  = 𝑀+ , it must also be
𝑀󸀠 = π‘Žπ‘Ÿ ⋅ ⋅ ⋅ π‘Ž1 |0 ⋅ ⋅ ⋅ 0(−1); that is, 𝑀 = 𝑀󸀠 , absurd.
Proposition 26. Consider ((1.1a) AND (2B) AND (3.2b)) ⇒
absurd.
(32)
2Aσ³°€
Absurd 3Aσ³°€
2B
2C
3Aσ³°€
3B
3.1b
3.2b
2D
3Aσ³°€
3B
3.1b
3.2b
3B
3.1b
3.2b
(33)
We observe now that by virtue of Proposition 5 when we
start from the condition (1B), we must consider two subtrees,
the first with root (1.1b) (which we denote by TREE(1.1b))
that is
Journal of Applied Mathematics
9
(1AσΈ€  ) AND (2D) AND (3.1b) ⇒ absurd by Θ1A and
Proposition 24
TREE (1.1b):
1.1b
2Aσ³°€
2B
σ³°€
Absurd 3A
2C
3Aσ³°€
3B
3.1b
3.2b
(1AσΈ€  ) AND (2D) AND (3.2b) ⇒ absurd by Θ2D and
Proposition 11
2D
3Aσ³°€
3B
3.1b
3.2b
3B
3.1b
3.2b
(1.1b) AND (2B) AND (3AσΈ€  ) ⇒ absurd by
Proposition 9 (i)
(34)
(1.1b) AND (2B) AND (3.1b) ⇒ absurd by
Proposition 9 (i)
and the second (which we denote by TREE(1.2b)) that is
TREE (1.2b):
(1.1b) AND (2B) AND (3.2b) ⇒ absurd by
Proposition 9 (i)
(1.1b) AND (2C) AND (3AσΈ€  ) ⇒ absurd by Θ3A and
by Proposition 21
1.2b
2Aσ³°€
2B
σ³°€
Absurd 3A
2C
3Aσ³°€
3B
3.1b
3.2b
(1.1b) AND (2C) AND (3.1b) ⇒ absurd by
Proposition 9 (iv)
2D
3Aσ³°€
3B
3.1b
3.2b
3B
3.1b
(1.1b) AND (2C) AND (3.2b) ⇒ absurd by
Proposition 9 (iii)
3.2b
(35)
(1.1b) AND (2D) AND (3.1b) ⇒ absurd by Θ2D and
by Proposition 9 (ii)
Let us note that the ramification
3B
(36)
3.1b
(1.1b) AND (2D) AND (3AσΈ€  ) ⇒ absurd by Θ2D and
by Proposition 9 (ii)
(1.1b) AND (2D) AND (3.2b) ⇒ absurd by Θ2D and
by Proposition 9 (ii)
3.2b
in each of the previous trees follows by Proposition 7.
We have then exactly 27 possible paths from the roots to
the leafs in all the previous trees (we exclude the paths that
lead to (2AσΈ€  ) because we already know that they lead to a
contradiction by Θ2A). In the following we write all these
27 paths and, for each, the related proposition that leads to
a contradiction together with some implication among the
Θ1A, Θ2D, and Θ3A. In the sequel, the node (3B) is omitted
for brevity because it is only an intermediate node and in the
paths that involve (3B) we use only the leafs (3.1b) and (3.2b)
to deduce the related contradictions:
(1.2b) AND (2B) AND (3AσΈ€  ) ⇒ (1.2b) AND (2D)
AND (3AσΈ€  ) (that we examine below) by Θ2D and by
Proposition 10
(1.2b) AND (2B) AND (3.1b) ⇒ (1.2b) AND (2D)
AND (3.1b) (that we examine below) by Θ2D and by
Proposition 10
(1.2b) AND (2B) AND (3.2b) ⇒ (1.2b) AND (2D)
AND (3.2b) (that we examine below) by Θ2D and by
Proposition 10
(1.2b) AND (2C) AND (3AσΈ€  ) ⇒ absurd by Θ3A and
by Proposition 27
(1AσΈ€  ) AND (2B) AND (3AσΈ€  ) ⇒ absurd by Θ1A, Θ3A,
and Proposition 25
(1.2b) AND (2C) AND (3.1b) ⇒ absurd by
Proposition 9 (v)
(1AσΈ€  ) AND (2B) AND (3.1b) ⇒ absurd by Θ1A and
Proposition 23
(1.2b) AND (2C) AND (3.2b) ⇒ absurd by
Proposition 9 (iii)
(1AσΈ€  ) AND (2B) AND (3.2b) ⇒ absurd by Θ1A and
Proposition 26
(1.2b) AND (2D) AND (3AσΈ€  ) ⇒ absurd by Θ2D, Θ3A,
and Proposition 28
(1AσΈ€  ) AND (2C) AND (3AσΈ€  ) ⇒ absurd by Θ1A, Θ3A,
and Proposition 20
(1.2b) AND (2D) AND (3.1b) ⇒ absurd by
Proposition 9 (v)
(1AσΈ€  ) AND (2C) AND (3.1b) ⇒ absurd by Θ1A and
Proposition 24
(1.2b) AND (2D) AND (3.2b) ⇒ absurd by Θ2D and
by Proposition 11
(1AσΈ€  ) AND (2C) AND (3.2b) ⇒ absurd by Θ1A and
Proposition 9 (iii)
(1AσΈ€  ) AND (2D) AND (3AσΈ€  ) ⇒ absurd by Θ2D, Θ3A,
and Proposition 28
This concludes the proof of Theorem 4. Below we draw
the Hasse diagram of the lattice 𝑆(5, 3, 2) by using the
evolution Rules 1, 2, and 3 starting with the minimum element
of this lattice, which is 00|123. We label a generic edge of the
10
Journal of Applied Mathematics
next diagram with the symbol π‘…π‘˜ if it leads to a production
that uses the Rule π‘˜, for π‘˜ ∈ {1, 2, 3}:
𝑅3
:
:
:
𝑅3
𝑅1
:
𝑅1
:
𝑅1
𝑅2
𝑅3
:
(37)
𝑅3
:
𝑅2
𝑅3
𝑅3
:
𝑅2
:
:
References
[1] D. K. Arrowsmith and C. M. Place, An Introduction to Dynamical Systems, Cambridge University Press, Cambridge, UK, 1990.
[2] J. L. G. Guirao, F. L. Pelayo, and J. C. Valverde, “Modeling the
dynamics of concurrent computing systems,” Computers &
Mathematics with Applications, vol. 61, no. 5, pp. 1402–1406,
2011.
[3] W. Foryś, J. L. G. Guirao, and P. Oprocha, “A dynamical model
of parallel computation on bi-infinite time-scale,” Journal of
Computational and Applied Mathematics, vol. 235, no. 7, pp.
1826–1832, 2011.
[4] J. A. Aledo, S. Martı́nez, and J. C. Valverde, “Parallel dynamical
systems over directed dependency graphs,” Applied Mathematics and Computation, vol. 219, no. 3, pp. 1114–1119, 2012.
[5] J. A. Aledo, S. Martinez, and J. C. Valverde, “Parallel discrete
dynamical systems on independent local functions,” Journal of
Computational and Applied Mathematics, vol. 237, no. 1, pp. 335–
339, 2013.
[6] J. A. Aledo, S. Martı́nez, F. L. Pelayo, and J. C. Valverde, “Parallel
discrete dynamical systems on maxterm and minterm Boolean
functions,” Mathematical and Computer Modelling, vol. 55, no.
3-4, pp. 666–671, 2012.
[7] C. Bisi and G. Chiaselotti, “A class of lattices and boolean
functions related to the Manickam-Miklös-Singhi conjecture,”
Advances in Geometry, vol. 13, no. 1, pp. 1–27, 2013.
[8] G. E. Andrews, “Euler’s ‘De Partitio numerorum’,” Bulletin of
American Mathematical Society, vol. 44, no. 4, pp. 561–573, 2007.
[9] W. J. Keith, “A bijective toolkit for signed partitions,” Annals of
Combinatorics, vol. 15, no. 1, pp. 95–117, 2011.
[10] G. Chiaselotti, G. Marino, and C. Nardi, “A minimum problem
for finite sets of real numbers with nonnegative sum,” Journal
of Applied Mathematics, vol. 2012, Article ID 847958, 15 pages,
2012.
[11] T. Brylawski, “The lattice of integer partitions,” Discrete Mathematics, vol. 6, pp. 201–219, 1973.
[12] K. J. Al-Agha and R. J. Greechie, “The involutory dimension of
involution posets,” Order, vol. 18, no. 4, pp. 323–337, 2002.
[13] K. Brenneman, R. Haas, and A. G. Helminck, “Implementing
an algorithm for the twisted involution poset for Weyl groups,”
in Proceedings of the 37th Southeastern International Conference
on Combinatorics, Graph Theory and Computing, vol. 182, pp.
137–144, 2006.
Advances in
Operations Research
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Advances in
Decision Sciences
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Mathematical Problems
in Engineering
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Journal of
Algebra
Hindawi Publishing Corporation
http://www.hindawi.com
Probability and Statistics
Volume 2014
The Scientific
World Journal
Hindawi Publishing Corporation
http://www.hindawi.com
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
International Journal of
Differential Equations
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Volume 2014
Submit your manuscripts at
http://www.hindawi.com
International Journal of
Advances in
Combinatorics
Hindawi Publishing Corporation
http://www.hindawi.com
Mathematical Physics
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Journal of
Complex Analysis
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
International
Journal of
Mathematics and
Mathematical
Sciences
Journal of
Hindawi Publishing Corporation
http://www.hindawi.com
Stochastic Analysis
Abstract and
Applied Analysis
Hindawi Publishing Corporation
http://www.hindawi.com
Hindawi Publishing Corporation
http://www.hindawi.com
International Journal of
Mathematics
Volume 2014
Volume 2014
Discrete Dynamics in
Nature and Society
Volume 2014
Volume 2014
Journal of
Journal of
Discrete Mathematics
Journal of
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Applied Mathematics
Journal of
Function Spaces
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Optimization
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Download