Let be a family of graphs, and a “host” graph. A spanning subgraph of is called - saturated in if contains no member of as a subgraph, but contains a member of for any edge . We let be the minimum number of edges in any graph which is -saturated in , where if contains no member of as a subgraph. Let be the -dimensional grid, with entries in each coordinate taken from , and the complete graph on vertices. Also let be the family of all subdivisions of a graph . There has been substantial previous work on extremal questions involving subdivisions of graphs, involving both and the Turan function , for or a complete bipartite graph. In this paper we study for the host graph , and , motivated by previous work on . Our main results are the following; 1) If at least one of or is odd with and , then 2) For even and , we have 3) For with even and , we have .
Suppose is a family
of graphs. We say that a graph is
–saturated
if no element of is a
subgraph of , but for any edge
in the complement of , contains a subgraph isomorphic to
some . When is a single graph
, and G satisfies the preceding
conditions, then we say that is
-saturated.
The well studied Turan function is the maximum number of edges in any -saturated graph on vertices. A natural dual to is the saturation
function , which
is the minimum number of edges in any -saturated graph on vertices. We write and for these two functions when
; that is, when
consists of the single
graph . The best known upper bound
on for an
arbitrary family is
given in [27].
A natural first case to study for the saturation function is , where is the complete graph on vertices. It was shown in [39] and later independently in [15] that , and that the upper bound is
uniquely realized by the join ; that is, the disjoint union of with a set of independent points, together with
all possible edges
joining the set of independent points with the vertices of the . At the opposite extreme in edge
density from for connected
graphs is the star , where
there is the following result.
Theorem 1.1. [27] The values of
are given as
follows:
If , then
, then .
Let be a fixed graph (and we
think of as the “host”), and
a given graph (where we think of
as the “forbidden” graph). A
subgraph of is called –saturated in (or just –saturated if
is understood) if is not a subgraph of , but for any edge the graph contains the subgraph . We then let and be the maximum and minimum
(respectively) number of edges in a graph which is -saturated in . If is not a subgraph of then we let . Observe
also that
and .
This saturation problem for general host graphs was to our knowledge first mentioned in
[15]. Erdos examined in [13]. In [15] the value of , and , was conjectured, where we specify
that the -set (resp. -set) of the occur in the -set (resp. -set) of the . The conjecture was confirmed
independently by Bollobás [6], [3], and Wessel [], [37], where they showed that . Let be the saturation function for the unordered version
of this problem; that is, where we allow the -set of to be either in the -set or the -set of , and the -set is in the opposite set of . In [33] it was conjectured that . In [20] it was shown that and that the conjecture holds in the case and .
Another host graph that has been considered is the complete
multipartite graph. The above saturation results for bipartite hosts
were generalized to -uniform
multipartite hypergraph hosts by Alon [2]. In this paper he proves a result on extremal
sets using methods of multilinear algebra, and from this derives results
on saturation in multipartite hypergraphs. The bipartite host results
were also generalized by Pikhurko in his Ph.D. thesis at Cambridge [34]. Now let be the complete multipartite
graph on partite sets, each of
size . In [18] the values of were determined for
all for large enough , and also determined for for all .
For any graph , let denote the set of all graphs that
can be obtained from by inserting
any number of points of degree 2 along any set of edges of , and we call any such graph a
subdivision of .
Some also call this set of graphs , and any member of a topological. Slightly abusing notation and
for brevity, we sometimes refer to an arbitrary member of as an or just by the symbol .
There has been substantial research concerning subdivisions of graphs
in the context of both the saturation function and the Turan function . A direct influence on our work
here was the paper [17],
which we discuss below. In [29] the set of possible values for -saturated saturated graphs (where is the cycle on points) was determined. Next we mention
results related to the Turan number for subdivision families. In [5] the authors investigate , where is obtained from a cycle by joining a
point on the cycle with edges only to points off the cycle, or only to
points on the cycle, and again
is the family of all subdivisions of such an . We mention also the deep papers [24], [23], and [12], among others, exploring where is the complete bipartite graph in some
cases. One of the goals in this research is to find, for a given graph
the exponent for which . Another goal is
to show that for any such there
is a graph for which .
Moving closer to “gridlike” graphs, we mention work on saturation in
the hypercube, denoting by
the hypercube of dimension . In
[11] it was shown that , and conjectured that this bound was best
possible. In [25] this
conjecture was disproved, where it was shown that for every fixed , there exists a -saturated subgraph of with edges. They also improved on
the earlier bound on by showing that and asked for which is it true that . It was
shown in [32] that this holds
for every , specifically,
that .
In a related paper [31] to
this one, the authors considered as a host the multidimensional grid
; that is, the -fold cartesian product of paths on points as the host graph, where the
guest graph is the star . In
that paper the following results for the function were proved.
(This result for was implicit in results of [28].)
For and even, , where is a constant.
For and odd, , where is a constant.
(This also
holds for .)
In this paper we continue the study of saturation in the host . Let be the family of all graphs
which are subdivisions of . In
this paper we obtain results on . Definitions and
the precise statement of results follow in the next subsection.
It is useful to mention the connection of the saturation function
to applications in the
area of bootstrap percolation, where we have drawn
on [8] and [30] for the overview which follows.
In that area the general idea is to analyze diffusion or infection
processes which spread through the vertices or edges of a graph. As
such, this models in this area are useful in describing magnetic
materials, fluid flow in rocks, computer storage systems, and spreading
of rumors. So this area has been of interest to physicists, computer
scientists, and sociologists; see [1] and [35].
In the vertex version, we begin with a starter set of vertices in some graph , and then build a sequence of sets
, , where for , according to some rule. Letting
, one question of interest is whether ; that is,
whether the process diffuses to all vertices in . In this case we say that percolates, or that is a percolating
set for .
A popular model of bootstrap percolation is –neighbor bootstrap
percolation, where the diffusion process is given by . Thus a vertex joins set if at least of its neighbors are already in ; that is these neighbors have
already been infected by time .
One may choose randomly (see
[9]); that is, each vertex
is initially infected with probability independently of all other vertices. We
then consider the event , and define the critical percolation
probability by
The class of -dimensional grids
is of interest as a
possible graph for this work.
Among the results here we mention the work of Holroyd [22] who proved . Holroyd introduced a
function (which we
omit here), which was used by Balogh [4] in proving the sharp threshold for all
dimensions given by
where
is the -times iterated logarithm.
Another model considers a diffusion process through edges called
–bootstrap
percolation. As notation, for any set of edges in some graph , let denote the subgraph of of induced by . Now let and be graphs, and such that is not a subgraph of . We say that is weakly -saturated in if there is an ordering of the edges
in such that for the
edge sets , and generally , , the following holds. For each there is a subgraph contained in , but not contained in . That is, the addition of edge
creates a copy of in not present in . The research problem, proposed
by Bollobás in [7], is
then to find a minimum size weakly -saturated set in , denoted by .
In an equivalent model for -bootstrap percolation, starting again
with some , we
build a sequence of edge sets of according to the following rule:
If for some , then we say that our “starter”
set is an -percolating set in . Here we add entire sets of edges to
our percolating process at a time. Note that for any we have that is weakly -saturated in if and only if is -percolating in . Clearly the minimum size of an -percolating set in is also . In this second model one can
also ask for the maximum time ,
over all starting sets , until
the percolation process stabilizes; that is until . We mention [8] and [30] among others for work on this topic.
For the connection between the saturation function studied in this paper and
percolation, observe that . As proof, take any edge set realizing . Then is in fact weakly saturated in
. This is because by definition
, and for any
we have . Thus under
any ordering ,
of the edges in the
sets given by , and for gives a process witnessing that
is weakly -saturated in ; that is a process for which , where . Thus .
Bollobás conjectured that we have equality for the -percolation process in ; that is , the last equality having
already been mentioned in our introduction. This conjecture was verified
independently in [2], [26], and [19].
Thus the upper bounds and exact results on provided in this
paper also act as upper bounds for ; that is, for
the minimum size of starter set in the -bootstrap percolation on . More generally one can use
upper bounds on to give
upper bounds on the minimum size of a starter which percolates in the -percolation process on .
Finally we mention the dynamic survey [16] which gives a broad and detailed coverage of
the area of saturation in graphs and hypergraphs. Another useful survey
is given by Gould in [21],
among other works on saturation by this author.
1.2. Definitions and saturation
in grids
We begin with some definitions, starting with the multidimensional
grid for positive
integers and . It has vertex set an integer with , so vertices are -tuples with ’th coordinate being an integer in . The edge set is given by So is an edge in precisely when and disagree in exactly one coordinate, and
in that coordinate they differ by . A straightforward induction shows that
.
For any subgraph of , we let be the subgraph of induced by points of with ’th coordinate equal to , . For an edge , we say that is a -dimensional edge if and disagree by in their ’th coordinates, and hence that for . For a graph and subgraph , we say that an edge is a nonedge of
in , and we let be the subgraph of induced by the set of nonedges of in .
Next, given graphs and we let , called the cartesian
product of and , be defined by and and , or and Note that can be obtained from by replacing each vertex of by its own copy, call it , of , and joining two such copies and by a matching joining
corresponding points in these two copies precisely when . Symmetrically, can also be obtained by doing
the similar replacement of vertices of by copies of . Thus is just the -fold cartesian product . In our illustrations of and its subgraphs, rows
are drawn from top to bottom in increasing row number. Also columns are
drawn left to right in increasing column number. So point is in the
’th row from the top and the ’th column from the left.
We now turn to subdivisions of complete graphs as the forbidden graph
for the saturation function. Clearly any member of is a cycle, so as realized by a
spanning tree of . Next, it is
straightforward to show that any minor is a member of (see also Proposition 1.7.2 in
[13] for a more general
observation). It is also shown in [13] (Proposition 8.3.1) that any edge maximal
graph without a minor must be
a -tree, so it follows that .
An exact result for and upper bounds for were given in
the following result. This result motivated our work on which follows
that result:
We use standard graph theoretic notation, as may be found for example
in the texts [38] or [10]. There will be additional
notation introduced when needed throughout the paper.
2. -saturation in dimension 2
The following Lemma gives the lower bound for our exact result in
dimension :
Lemma 2.1.
Given a connected host
graph , we have .
Proof. Let be an
-saturated subgraph of
. In particular is spanning in and contains no . We claim that must be connected. Suppose not and let
and be distinct connected components of
. Since is connected there must exist an edge
having exactly one
end in one (possibly in each) of and . So is a cut edge of . Since , this copy of must contain since contains no . Now , being a cutedge of is contained in no cycle of , and therefore in no cycle of . Thus is a cutedge of . But any must be edge connected, contradicting being a cut edge of this .
Thus we know that is a
connected spanning subgraph of ,
so . Take
any . Suppose first
that , so that
is a spanning tree of . Then and contains only one cycle, contradicting
that its subgraph contains
cycles. Next suppose that , so that is a tree plus an edge. Then , being a tree plus two edges, still
contains fewer than cycles, again
contradicting that contains cycles.
Thus , as
required.
We will need the following obvious remark for future reference:
Observation 2.2. An consists of a cycle and a vertex , together with three vertex
disjoint paths (apart from )
joining to . We will call such a set of three paths independent.
Theorem 2.3.
If at least one of or is odd, , , then .
Proof. By Lemma 2.1 it suffices to prove the upper
bound . We will construct an -saturated subgraph from
which the upper bound follows. An example is given in Figure 1.
Figure 1 Graph realizing
The graph may be
obtained from ,
odd, by removing the following
sets of edges:
Remove the set of edges ; that is,
remove the first edge of row ,
.
Remove the set of edges even, ; that is, remove the
last edge in each row with even.
Remove the set of edges odd, ; that is, remove the
second edge in each row with
odd, .
Remove the set of edges ; that is, remove all edges in columns , .
We now verify that is a
-saturated subgraph of
with odd. The reader can use Figure 1
illustrating when reading
the following case arguments.
First we note that
contains no , since contains exactly cycles, while an contains cycles.
It remains to show that for any nonedge of
, contains a cycle and vertex as described in Observation 2.2. As notation, when and are two vertices of along the same row or column of
, we write to mean the path in along that row or
column joining joining and . Let be the length cycle of having the vertices, listed in a
traversal of , , and let be the length cycle having the vertices . We consider the following
representative cases.
Suppose first that is a
horizontal nonedge; that is, . Consider first the case , even. Then let , and let . Then the three independent
paths are ,
, and . The case for odd is similar, using the same and . The final kind of horizontal nonedge is of the form
, for some . Then let , , and the three paths are , , and .
Second, suppose is a vertical
nonedge, starting with the case , with even,
, and . Then let , , and the three paths are , , and If ,
then take and the cycle
. Then we have three independent paths in given by , , and .
The other type of vertical nonedge is , with is
odd. Consider first the special case , so ,
. Then take the cycle
. Now take , and we have the three paths , , and . For the general case
with (still with
, odd) take the cycle to be , and the three paths are , , and .
3. -saturation in dimension 3
In this section we give a construction that realizes upper bounds for
, even and . It provides the basic intuition
behind our general construction realizing upper bounds for for arbitrary
dimension that we give in
the next section. We refer to the four points of degree in an as junction points of
that , and to the six
internally disjoint paths in this joining pairs of junction points
as junction paths.
For any subset , , we
let be defined by ; that is is obtained by appending as the ’st coordinate and leaving the first
coordinates unchanged in each
point of . For example if , then . Similarly if
is a subgraph of , then is the subgraph of induced by the vertex set
. So clearly under the natural
isomorphism over all .
We begin by constructing a subgraph for even:
3.1. Construction of
subgraph ,
even
Let
Let be the subgraph
of induced by the
boundary cycle of ; that
is, is the graph induced by
the set of edges or or .
For every vertex and , ,
define the path by;
if is even, then , and
if is odd, then .
For , let .
Let
Let .
End of construction of .
Each of the two rectangles in Figure 2 consisting
of straight horizontal or vertical edges is a copy of . The curved lines are used in the
construction of which follows, and can be ignored for now. Apart from
the boundary of , the
horizontal straight lines represent paths . For example, (suppressing third
coordinates) we have and . We will sometimes refer to as a square. Since
is connected and contains a
unique cycle (induced by its boundary), we have . Trivially contains no since contains exactly one cycle, we now
generalize to obtain our
construction of , an -saturated subgraph of . We begin with an informal
description.
Figure 2 The graph ; horizontal arcs are paths in
3.1.1. The cross sections
and
are
defined as copies of ; that is
we let and
. We link
and by a pair of length paths and in , where (resp. ) is the shortest path in joining to (resp. joining to ). In fact all edges
of of and of are 3-dimensional. Next for each
we will define a path
of 3-dimensional edges such
that the paths are
vertex disjoint and each path is pendant on exactly one point of
. We the let
.
The role of the paths
is to span the gap of vertices lying between and ; that is, those with , in such a way as to
make a connected spanning
subgraph of while also
ensuring that is -saturated. In this respect the
collection of paths
plays the role in that was
played by the set of paths , each path
being pendant exactly one point in .
Letting , and , we then have the decomposition , in analogy with
the formula for . We may view
as a “skeleton” of consisting of two cycles and
attachment paths and . The subgraph fills in the gap isolated points in
, with
doing the filling using
dimension 2 edges and
using dimension 3 edges.
3.2. Construction of
subgraph ,
even
Let
Let the two cross sections and of be corresponding copies of . That is, we let
and .
Define a ‘left attachment path’ joining and , and a ‘right attachment path’
joining and by
, and
.
Let (resp. ) be the unique cycle of (resp. ). Then let
.
(Construction of ). Take
the unique bipartition of for which .
For every vertex and , where define the paths and by;
if , then
,
if , then
.
For ,
let .
Recalling the subgraph , let
Define sets by
Let .
End of construction of .
We illustrate the graph in
Figure 2.
In that Figure the two large rectangles stand for and , each a copy of , as described after
the construction of . The
straight horizontal lines represent rows in and . The curved horizontal arcs
represent paths in . The darkened points at the ends of
the arcs are the endpoints of such a , and are either the pair and if , or the pair and if . The open circled points are
those immediately preceding or following a path along a nonedge of dimension . For example, the path begins at the
darkened point , passes
through the points , in order, and ends at the
darkened point . Then
is followed by the
open circled point joined to by a -dimensional nonedge of in .
We can now prove -saturation for . We will say that is a pendant subgraph in (or of ) if contains a cutpoint of such that is one of the connected components of . Thus is a -component of , and in this case we will also say that
is pendant in at.
Theorem 3.1.
Suppose with even. Then is an -saturated subgraph of .
Proof. First we show that contains no , and for this suppose the
contrary for contradiction. We let be a copy of an in . Let and be the unique cycle in and respectively.
Observe that the subgraph
of induced by is a forest, each component of which
is pendant on .
Indeed, each such component is
either
a comb with base ; that is, contains some path , and is a disjoint union of paths – these paths being of
the form for ), or
a path for some in row or row of either or .
Hence no edge of belongs
to a cycle of .
So by the above paragraph, is
a subgraph of . Now the are vertex disjoint cycles, and
and are vertex disjoint paths each
having exactly one point in common with each . Therefore contains exactly six cycles. But an
contains seven cycles, a
contradiction.
Next we show that for any nonedge of in , we have . For this, it
suffices to construct a cycle , a point , and three
independent paths, in accordance with Observation 2.2.
We will need some notation. We let indicate a path from vertex to vertex , where the symbol refers to the path used. For example,
can be the row number of column
number within either of the copies or of , or a path , or one of the paths or . We then indicate the concatenation
of such paths by writing , resulting in a path from to . The symbol will refer to an edge in joining vertices and , and we will abbreviate (resp. ) by (resp. ).
The nonedge falls into
one of the following categories:
and both belong to , or and both belong to ,
and , where and are at vertical or horizontal
displacement in ; that is, and , or and , .
for some , and . This possibility reduces to and for , or and for . That is, is the -dimensional nonedge preceding or
following one of the paths .
, and
.
Consider first category (a). Since , we
consider just the case . Here we pick some representative examples of such
nonedges . Consider first the
subcase where are in successive
rows of , neither row being
on the boundary of . So we
have and with . We shall first assume
that is even, so that row is the path , while
row is the path .
Then our cycle
is:
We then have the point , together with the three independent paths in given by
intersecting
at .
The construction of and three
independent paths is very similar if is odd, where this time the new intersects at , the new intersects at , while the new is the same as the old one so has
-intersection at .
Next consider the subcase (still in category (a)) where one of the
successive rows containing or
lies on the boundary of , at first taking rows and . So is the nonedge with and , where . Since is odd, row of is the path ,
while is a
nonedge of . This case is
illustrated in Figure 3 where we have outlined the cycle in
solid lines and three independent paths from to this cycle in dashed lines
which are numbered. The reader may wish to use this Figure while reading
the formal definition of the cycle and three paths which follow.
Figure 3 showing ; cycle in solid, three
independent paths from to
that cycle dashed and numbered
We have the cycle , point ,
and three independent
paths as follows.
A very similar argument applies when the boundary row (among the two
successive rows) is row . The
nonedge is then for some . The cycle becomes
, and we take together with three
independent analogous to the
ones given in the preceding paragraph,.
We remark that the assumption
even is necessary in the preceding subcase. If was odd, then row of consists of the edges . In that case, for the nonedge there is no subgraph in .
To complete category (a), consider the case where and are in the same row of . So for , is the nonedge if is odd, or the nonedge if is even. The proofs being very similar,
we suppose that the nonedge is with
odd. Then we have the cycle given by
.
Taking the point , we obtain the three independent paths
Consider now category (b). Suppose first that , . We first treat the case
where is even, with , so . The arguments are
similar, with minor changes, in the other combinations of possibilities
for the parity of and the partite
set membership of and . We can take since if then both , a case we’ve already
treated above. We have
since and are both even. Since and , we have and also
.
We then consider the cycle
Taking we obtain
the three independent
paths
As a special case in b) not covered above, take illustrating the case odd as well as row being on the boundary . So , ,
and is a
nonedge of . Therefore the
path used in the immediately preceding is no longer available for our construction. Instead we
construct a cycle , take the point
, and form three
independent paths as follows.
Still in category (b), the other possibility is , . Again we assume , and , and is even, other combinations of cases
yielding similar constructions. Let be the path Then we have the
cycle
We take
the point , and
obtain the three independent paths and
Next consider category (c). The parity of determines the nature of row according to step in the construction of . Also the parity of determines which partite set ( or ) contains , and hence the nature of the path
according to step
in the construction of . Specifically, if is even, then for we have that begins with the edge , while if is odd, then begins with the edge .
We first assume is even and
that is even for concreteness,
so that also is even. Later we
also consider odd and odd, with the other two combinations
of parities for and are omitted since they are similar to
the combinations considered here. So with and even we know that begins with the edge and ends with the point
. So our nonedge is , where and . We also assume that , explaining the last assumption
and treating after the
construction which follows. With
even, row in is , and the
same for row of except with third coordinate
. We then get an , starting with
the cycle is given by
Now consider the point , and we obtain the three independent paths
The assumption ,
together with even, allows for
the existence of the path used in , while does not allow this. This is because with the successive pair of that path would be a
nonedge (since is even).
We mention in brief the case with even, still in
the case and even. The cycle is given by
We take and
form the three independent
paths as follows. We let
(resp. ) be the (resp. ) path along the boundary
of starting with row (resp. column ). Then let be the path that uses , row , and column of .
Still under category (c) consider the possibility odd, and this time we take the case
odd. So this case implies even, and hence includes the case (though now with odd). With odd, row in is . Since
is odd our nonedge satisfies , and , and is the -dimensional edge preceding the path
.
Then in we have a cycle
, the point and three independent
paths as follows:
or or
Finally consider category (d). Here by symmetry we can take , say with , where since with (so that ) we have ,
so there is no vertex ,
, with a nonedge of . It follows that for some neighbor of in . So or , and since the constructions are
similar in these two cases, we consider here only , so . Since , then , so by construction is the path . Now
in we construct a cycle
, taking the point , and three independent
paths as follows:
It follows that for every nonedge of we have . This, together
with containing no , completes the proof that is -saturated in .
Corollary 3.2. For even with , we have
Proof. The lower bound comes from Lemma 2.1. For
the upper bound, it suffices to show that by Theorem 3.1. The number of edges of whose joint removal from leaves a spanning tree of is , where . Such a set of edges
can consist of one edge from the unique cycle of , a second edge from the unique
cycle of , and a third edge
from either of the paths or
. Therefore .
4. -saturation in dimension
In this subsection we construct an -saturated subgraph of for .
We iterate the previously defined operation. For any vertex , and any fixed -tuple , we let . Then for any subgraph
, we define the
subgraph
as the subgraph of
induced by the vertex set .
We generalize to a
subgraph ,
, which serves as a skeleton
of , and is defined
inductively as follows. Given , we let , where are paths linking the
two layers given by , and . Let be the
four endpoints of these two paths; that is, . Note then that contains squares, linked inductively by
sets of paths , at
dimensions . We will see
that the skeleton is -free.
But it is not -saturated
because of the gap of points at dimension
lying between and ; that is, points with , these being are
isolated in . Here we note
the role of the set as an
exception. That is, for the endpoints of and , and respectively, there
is no such gap, since every point on the shortest path joining each of
these pairs is already included in on either the path or . For all other pairs , differing only in
their ’th coordinate with values
and , this gap will be filled
by a path generalizing . We describe informally below, before
presenting its formal construction, as part of the inductive
construction of , .
First we will need additional notation. Take a point , and let be its
corresponding point in . Now define two paths
and
by
and
For each such pair we will choose
to be one of the paths ,
the choice based on a bipartition of . We then let . Now let (resp. be whichever of
or is (resp. is not) on . For example, in Figure 2 illustrating , we have while , and while . So one of
is in while the other
is in . The paths “fill in” the gap in
of points in lying on the
shortest path in between
and , and are the analogues of the
paths in the construction of
.
We also let and . Recalling the operation defined earlier in this
section, we have isomorphic copies of the paths and in given by and
for ,
the latter defined by .
Iteratively for any -tuple
with entries from , we also have and similarly for . Further for any set of paths of the form or (over various and ), , we let .
As remarked above the paths fill in gap in at dimension . The paths are “lifted” to isomorphic copies
of themselves in by iterated applications of the operation. Now will be the collection of these
lifted paths , , defined inductively by . The collection of paths “fills in” the previously
described gaps of isolated points in , in all dimensions , . With this inductive definition of we finally take .
Throughout the remainder of this section we shall occasionally use
the symbol (resp. ) as an abbreviation for (resp. ).
4.1. Construction
of -saturated subgraph
for
(Initialization) Construct subgraphs (given in section
3 ), , and given by (with
defined in section 3),
as defined in step 4 of the
construction of .
For assume subgraphs
, sets
of paths , and sets of attachment points have been
constructed.
Define attachment paths and attachment points in by
the left attachment point , and right attachment point
attachment paths given by
left attachment path , and
right attachment path
Define sets of attachment points and sets of attachment paths in
by
(set of left
attachment points in )
(set of right
attachment points in )
(set of left
attachment paths in )
(set of right
attachment paths in )
Let , and take the unique bipartition of for which .
(Construction of paths ). Consider any corresponding
pair and (that is, for and ), satisfying
,
Noting that , define paths by
If , then
If , then
.
(Construction of ).
where .
(Construction of ).
Let .
(End of construction).
In Figure 4 we illustrate the inductive
construction of , omitting
the set of paths in order to
simplify the illustration. The two copies of (namely and ) are encircled by large by
dotted ovals, and joined by the paths and . The eight boxes each represent a
copy of . The four
horizontally close pairs of these boxes, together with paths and , each form a copy of , where is a string of length over the letters that depends on the pair. The
top two copies of (i.e.
within ), together with
the paths and , form the copy of , while the bottom
two copies of together with
the paths and for the copy of .
Figure 4 The inductive construction of , omitting paths in this illustration. Also shown
are concatenations of left and right attachment paths (bold or dashed)
for Lemma 4.9
The and notation used above are iterated
in the natural way. For example, if and are strings over the letters and , then . Also we
have , ,
and where ,
are each or .
We can decompose as
follows. Let , we have .
For the first component on the right we have ,
giving us a copy of . We now
show that the second component is also isomorphic to a copy of , which reduces to showing . First
note that by step of the
construction we have . Therefore
we have
Similarly
We can now construct an isomorphism which is a
reflection in the ’th coordinate.
Take , so
with . Then
define . Observe that since leaves the first coordinates undisturbed, we have
and . By this reflection action in the ’th coordinate, we see that reflects the paths and about their midpoints, so that
and . By the same reflection
we see that for any path , , and that
is then a bijection between and sending each path to its reflection . It follows that
using the decompositions of and in the preceding paragraph.
The fact that is a graph
isomorphism is straightforward from its reflection property. We argue
just a few typical cases here. For any , write ,
with . First take an edge . Then and , the case being similar, and
noting that . Then
and . This shows that
if then , and the same
statement with in place of
. It also shows that if , say with ,
then . Thus in the case
we have . As another case suppose ,
so and and .
Then . A similar argument
shows that preserves edges in
.
We now outline briefly how
preserves nonedges of in . For
brevity let
Nearly the identical proof given for edges shows that preserves nonedges of in lying in , based again on the reflection of
. We consider in detail one
different case; that is, where a nonedge of satisfies and . In this case, is the -dimensional nonedge immediately
following or preceding a path . We can take by symmetry and ,
with . Then which is a nonedge in . The remaining cases of
nonedges satisfy ,
whose details we omit. The proof that is a nonedge relies on the
nonexistence of any edge in
with one end in each of two distinct paths and , or one end in and the other in some
path .
So our decomposition shows that decomposes into two copies of
(the subgraphs and ), joined by paths and , together with the subgraph . Further, is a vertex disjoint collection
of paths of the form ,
where each such path is pendant on exactly one of or . This is illustrated in Figure
7.
As examples of the and
sets, we gave and in step of the construction of in the preceding subsection. We
also have and
Our goal now is to show that is an -saturated subgraph of . As our first step, we will
show that contains no . For this, we first reduce to
showing that the subgraph of
contains no . We let be the subgraph of
induced by . We call the vertices of connection points of , and we call any path in a connection path of . We examine for in the Lemma which follows.
Lemma 4.1. For , is a forest, each component tree
of which is pendant in on
.
Proof. Recall that we set and , and .
We proceed by induction on ,
starting with the base case .
We have already observed in the proof of Theorem 3.1 that
is a forest, each tree
component of which is pendant on either or , these being the unique cycles of
and respectively. So the Lemma holds
for .
Suppose the Lemma holds for ,
and consider . Let (resp. ) be the copy of in the subgraph (resp. ) of . Also let (resp. ) be the copy
of in (resp. ), so that .
We will use the decomposition
given after the construction of , where we showed that the
Applying induction to each of and , we see that and are forests pendant on and respectively. These two
forests are vertex disjoint since the last coordinate of any point in
the first is , while the last
coordinate of any point in the second is .
Consider then , the third
union component in from
step of the construction of . By construction is a vertex disjoint union of
paths . It suffices to show
for each path
that either
is itself a tree
component of pendant on
, or
is pendant in on some tree component .
In case (2),
is then a tree component of
inheriting the inductive property from of being pendant on .
Now path is pendant in
at either or (with the point corresponding to in as in step 6 of the
construction of ), and by
symmetry we can take take . Suppose first that , and by induction let be the tree component of containing . Then is pendant on , and thus case (2) is satisfied.
If , then by
construction . If
also for some tree
component of , then by
induction is pendant in
(and hence in ) at . Since also is also pendant at , case (2) is satisfied again. Still
with , if is in no nontrivial tree component of
, then clearly is pendant on at , and is therefore its own tree
component of . So case (1) is satisfied.
For the structure of ,
observe first that is a
disjoint union of paths of the form . We have previously observed in
the proof of Theorem 3.1 that is a forest each tree component of
which is a path or a comb; the latter being a tree
containing a path such that is a disjoint union of paths, and
of course a path being a special case of a comb. The path in these nonpath combs is a path for some . This can be seen in Figure 2, where the path plays the role
of and union of paths , and ,
together with form a tree
component which is a comb. For , the tree components are either trees containing a path
for for which each component of
is a comb, or just paths
paths (for example . For general , each tree component is a nested
sequence (in the sense above) of combs up to depth , though we do not use this fact.
In light of Lemma 4.1, for any , let be the
tree component of containing
. Then let be the point of
at which is pendant. As an example, we refer
back to Figure 2 and the tree , where is the path given in the above paragraph. This
is a tree component of the forest
, and is pendant in at the point . So for every vertex we have . In particular we
also have
for and generally for
each we have , .
Corollary 4.2. Take
If contains an , then this is a subgraph of .
For , there is a
unique path in from to .
Proof. For a), any vertex of an is contained in a cycle of that
. So by Lemma 4.1 no vertex of
of can be in
this cycle. Since , this is a
subgraph of .
Part b) follows immediately from Lemma 4.1 and the fact that is a tree.
The arguments in the rest of this section will be inductive, and to
facilitate these we introduce here some abbreviated notation for
referring to substructures of .
Since will be fixed, we
will abbreviate and to and respectively. So (resp. ) denotes the subgraph (resp. ) of , this being the subgraph of induced by points for which (resp. ). We let and be the subgraphs of
induced by points
satisfying and respectively, and similarly let
and be the subgraphs of
induced by points
satisfying and respectively. So altogether we
have , while
for .
The left and right attachment point and path structure is
naturally inherited by and
as copies of in , and their subgraphs etc. as copies of in . The coordinates in of such points are obtained by
taking the coordinates of the attachment point in (or ) and appending the appropriate
coordinate suffix in . So we
have
and for or we have and So we have and
Similarly the attachment paths and in appearing in the copy (resp. of will be written as and (resp. and ) as in step of the construction of .
The various left and right attachment paths, all of length , join pairs of these attachment
points. For example, joins
to , and joins to . Now let and be the left and right
attachment paths in , . Then joins to , and joins to .
Toward showing that is
-saturated in we begin by proving that contains no . Recall that the points of
degree in an are called junction
points or just junctions, and we call the six internally
disjoint paths in this
joining the pairs of junctions junction paths.
Theorem 4.3. For and even, the graph contains no .
Proof. By Corollary 4.2, it suffices to show that contains no . We proceed by induction on
, having proved the base case
in Theorem 3.1. Assume the Theorem is true for
, , and we prove it for . We begin with the following
claim.
Claim 4.4. If contains a copy of an , then not all four junction
points of can lie in , and not all four of these
points can lie in .
Proof. We proceed by induction. The claim is vacuously true
for since contains no . Now assume the claim true for
, , and consider now the clam for
. We write (resp. ) for the subgraph (resp. ), or .
Suppose to the contrary that all four junctions of lie in (symmetrically in ).
Subclaim 4.5. There is exactly one junction path
of leaving .
Proof. By the first inductive assumption that contains no , we have . So
.
Hence there is a junction path which leaves , enters , and then returns to . So contains the subpaths and , and hence there can be only one
such junction path. If not, then some other junction path leaving and
reentering must also
use and and yet be internally disjoint
from , a contradiction.
Subclaim 4.6. Not all four junctions of can lie in , and not all of them lie in
.
Proof. Assume to the contrary that all four junctions lie in
, with a symmetric
argument if they all lie in . With as in the statement of Subclaim 4.5, we
regard as leaving on to arrive at , passing through on some path from to , and finally returning to along . Now can take essentially one of the two
following forms:
a path
from to , followed by the path from to (illustrated in Figure 5), or
the path from
to followed by a path from to .
The choices of or lead to symmetric arguments later, and
we will address only choice A here. We note that the existence of in choice (A) and of in choice B is ensured by Lemma 4.9b, which follows and is
independent of this Theorem. Now
must finally return to
via either the path or
the path . If then would intersect itself at the endpoint
of , a contradiction.
So we have .
Then .
This is because cannot cross
itself and also cannot exit
through since as just shown
. So after
passing through , reenters at and continues with a subpath
going from
to one of the two junction
points, call it , of joined by . Letting be the other of these two
junctions, we can write , with
choice A above, as , where is the subpath of joining to . This is illustrated in Figure 5.
Figure 5 Replacing the path by
dotted path in the
proof of Theorem 4.3
Observation 4.7. Any junction path besides
satisfies .
Proof. Suppose to the contrary that some junction path leaves . Then must leave and then reenter by the contrary assumption
to Subclaim 4.6 that that all four junctions lie in . So must contain two of the of the three
entry/exit paths into . But already contains two of these paths,
contradicting that and are internally disjoint.
We can now complete the proof of Subclaim 4.6. We still refer
to the junction path of Subclaim
4.5.
We “splice out” the subpath from , and replace it by the dashed subpath
contained in , as illustrated in Figure 5 and described as follows. Take any path
in with ends and , noting that such a exists by Lemma 4.9 which
is proved after and independently of this Theorem. Now replace in by the path , observing that and have the same two
endpoints and . So we obtain the new path ,
so .
We claim that can be taken as
a junction path replacing in
. First, joins the same two vertices and that does. Note also that the subpath of has no internal intersection with
other junction paths of since
such intersections, apart from possibly , are disallowed by the
Observation above, and for
disallowed by the original internal disjointness of with other junction paths. The other
three subpaths of , namely , , and , also have no such internal
intersections since these were already subpaths of which had no such intersections.
Therefore is indeed a path which
is internally disjoint from the other junction paths of , and hence can act as a junction path
in replacing . Since was the only junction path leaving
by Subclaim 4.5,
this replacement results in a copy of lying entirely in , contradicting
the inductive hypothesis of the Theorem that contains no copy of . So Subclaim 4.6 is proved. (end
of proof of Subclaim 4.6).
We can now complete the proof of the Claim, still working with the
contrary assumption that all four junctions of lie in . By Subclaim 4.6 we need only
consider the two nontrivial partitions of the junctions among and .
Suppose first that each of and contain two junctions. Then
there must be at least four internally disjoint junction paths exiting
and ultimately
entering . Now any path
exiting and eventually
entering must have as
a subpath exactly one of the three exit paths from in ; these being , and . By the Pigeon Hole principle
two of these junction paths share one of these exit paths, contradicting
their internal disjointness.
So we can assume by symmetry that contains a single junction
while contains the other three
junctions of . Then there must be
three internally disjoint junction paths having only vertex in common leaving and eventually entering
. Hence the three paths
, and are distributed in a one to
one way as subpaths of the three internally disjoint junction paths from
. One of these junction paths,
call it , contains , passes through , and must enter along the subpath . A second such junction path,
call it , enters along the subpath . But then , a
contradiction since these three paths can have only vertex in common.
Returning to the proof of the theorem, assume to the contrary that
contains a copy of an . By the preceding claim each of
and contains at least one
junction of . So there must be at
least three junction paths of
joining junctions in to
junctions in . But each of
these three paths contains exactly one of the two paths or , the latter being the only exit
paths from in . By the Pigeon Hole principle
this contradicts the internal disjointness of the three junction paths,
completing the proof of the theorem.
To complete the proof that
is -saturated in , it remains to prove the
nonedge addition property. To that end, we will need a Lemma on the path
structure in .
Definition 4.8. Let be a square, its cycle, and . Let be the point on closest to (so that if ). We let (resp. ) be the shortest path in
from to (resp. ). We illustrate , , and in Figure 6.
Figure 6 Paths in the square containing from the right and left attachment
points of
Lemma 4.9. For , let be a square in , and the unique cycle of this :
There exists a path from to
(resp. to ) which is a concatenation of
left attachment paths (resp. right attachment paths).
There is a path from to which is a concatenation of
right attachment paths (paths in ) from to , followed by a path in along from to , followed by a concatenation of
left attachment paths (paths in ) from to .
Let , . Then
(resp. ) passes through and avoids (resp. ).
With as in c), there is a
path in along from (resp. ) to intersecting (resp. ) only at .
Proof. In this proof we will refer to subgraphs or in , for suffixes or of length or (resp.) over the letters and , by their isomorphic copies or for brevity. These suffixes locate
the copy of or within , but that location is not important
for this proof.
Starting with part a) we prove the statement for the path from to , where the proof for the path
from to is identical with the obvious
changes. We proceed by induction on . In the base case , contains just the two squares and , each isomorphic to . Noting that , the required path
is then given by joining
and in the construction of .
For the inductive step, assume the theorem true for and consider with its two subgraphs and , each isomorphic to . We have or . If , again
noting that we obtain the required concatenation of left
attachment paths , for some
sequence of integers , by applying the inductive hypothesis.
Now suppose . By induction there is a path consisting of a
concatenation of left attachment paths joining to . Recall that by construction of
, the path joins to . Then we obtain the
concatenation of
left attachment paths joining to , completing the inductive
step.
For part b), again we proceed by induction. In the base case , for the required path we take the path
from to
around the cycle of
, followed by the path
joining to . Here the right attachment
path in the statement can be taken as the trivial one consisting of just
the point itself.
For the inductive step we assume the theorem true for . If , then apply
induction to find the required path in from to passing through as required. Now concatenate with the left attachment path joining to given in the construction
of to obtain the required
concatenated path joining
to for part b). If , then we begin
with the path joining and given in the construction
of . Now apply induction to
get a path in joining and and passing
through . Then the
concatenation gives the
required path in joining
to for part b).
An illustration of the concatenations of paths in parts a) and b) of
the Lemma is given in Figure 4, these
concatenations shown in bold and dashed curves. An accompanying
description is given immediately following the proof of this Lemma.
For parts c) and d), we refer the reader to Figure 6 which illustrates the proofs which
follow.
Consider part c). If ,
then is on a pendant path from
, so both and must contain . Also consists of this pendant path
followed by the shorter of the two paths on from to . Therefore avoids , and avoids . If , then and the claims on and are then obvious for the same
reasons.
Consider part d). By part c) and the definition of , the shortest path from to (which runs along ) is the path required in part d). The
proof for is just obtained
by the obvious interchange of with , and of with .
Figure 4 illustrates concatenations of
left and right attachment paths (in bold and dashed curves) claimed to
exist in parts a) and b) of this Lemma. For a concatenation of left
attachment paths from to
( as in part a) ) observe
the sequence of paths starting at the upper left corner which is the
vertex . One such
concatenation begins with the left attachment path followed by a an
inductively constructed concatenation of such paths (indicated by the
oblique line) within the upper and second from left copy of ending at the vertex of the within that copy of in the figure. Another example
again begins at , and
proceeds along the left attachment paths in the order and finally along a an inductively constructed concatenation
of left attachment paths (again shown as an oblique line) within the
lower and rightmost copy of , leading to . Finally observe the
concatenation of right and left attachment paths going from (the corner point at lower
right) to (the corner
point at upper left) as in part b) of the Lemma. This begins with the
succession of two right attachment paths followed by the
inductively constructed concatenation of right attachment paths within
the upper second from left copy of , leading to (all in dashed curves), followed
by the cycle around , followed
by the reverse of the concatenation of the first two left attachment
paths given in the first example, leading to .
In our next theorem we show that the edge addition property holds in
; that is, that for any
nonedge of in we have . The proof will
be by induction on , and to
facilitate that induction we need the following definitions and
notation.
We continue with the notation and its concatenation used
previously in the analysis of . We extend here the meaning of
symbol to include not only the
name of a path joining vertices
and , but also the name of a Lemma
or construction justifying the existence of such a path. So can take values to indicate Lemma 4.9 parts respectively, or to indicate Corollary 4.2 parts respectively, or just the name of
some path – usually some attachment path such as , , or .
Theorem 4.10. For , even and any nonedge of in , contains an .
Proof. We proceed by induction. The theorem has been proved
for in Theorem 3.1. Assume the theorem true for
, , and we prove it for . With the notation given just
before this theorem, we have the following decompositions:
These decompositions are illustrated in Figure 7. On the left you see
the decomposition of into
four copies of , the top two
being and contained in , and the bottom two being and contained in , together with various attachment
paths. The attachment paths
and link with , while the attachment paths and link and , and and link and . Those connection paths in lying outside the four copies of
are excluded for
simplicity, but are included in the right Figure 8. Figure 8 shows
paths of the form running
between and , and paths of the form running between and together with paths running between and .
Figure 7 excluding
Figure 8 containing and
Let be any nonedge of
in . If both or both , then we are done by the
inductive hypothesis as applied to or since . Further we
cannot have and since then , while at the
unique coordinate where and disagree the difference must be , a contradiction. Thus either or exactly one of lies in .
Let the nonedge be -dimensional. We argue according to the
value of . In the body of this
proof we treat only the case , and provide the similar details for the case in Appendix II.
The assumption ,
together with being a nonedge,
implies that is the unique
nonedge of dimension following
or preceding a -path, call it
, of dimension . By symmetry we can take and , so that
. So
in this case exactly one of
lies in , this being
. The situation so far is
illustrated in Figure 9, where additional aspects of that Figure
will be explained soon. We have two possibilities; either or . For convenience,
set , so that in each case we have .
Figure 9 An in when , cycle in solid lines, three
independent paths from in
dashed lines
Assume first that . Since we have or
, and by symmetry we can
suppose that .
We now show that . Now lies on
some -path of dimension , call it , with since . This is portrayed as the
horizontal line through in
Figure 9, of which the bolded part is the subpath
joining and . Similarly there is a -path of dimension , call it passing through , portrayed as the horizontal line
through in Figure 9. But and are corresponding points in and (resp.). So we have by construction.
Since
we get , so
and it follows that .
Recall that and are in by definition, so each of these
points is either in a square
or in some attachment path or
, . For the bound , observe first that since by construction there are
points for
which is defined. Also
since .
In the constructions which follow we will suppose that each of and is in its own square, and later
we treat briefly the case where each of and is in some attachment path
or by a small modification. Given the
reduction in the preceding paragraph, we then have and .
We then obtain an , as illustrated in Figure 9 and described
formally as follows. Define a cycle containing ,
indicated in bold lines in Figure 9, as follows.
We describe here in prose the use of the arrow notation above to
describe , to further familiarize
the reader with this notation. Here starting with the edge , we follow the connection path joining and
in given by Corollary 4.2b, followed by the
path given in Lemma 4.9c.
This then is followed by a succession of left attachment paths all
guaranteed to exist either by Lemma 4.9a or
the construction of and . These are, in the order they
appear on , given by , , , . These are
followed by the path justified by Lemma 4.9c, and then the cycle is
completed by the connection path given in Corollary 4.2b, the latter path
passing through and as shown in the Figure.
We now take the point ,
knowing that since
by construction never enters
. The three independent paths required to complete the
construction of an , indicated in numbered dashed curves and lines in
Figure 9, are as follows.
We give the proof of independence of the paths , in Appendix I. For the remainder of this section we omit
such independence proofs as they are very similar to the one given
there.
We now indicate briefly the adjustment to the constructions above
that will cover the case where each of and is on some attachment path.
Suppose for example that . This is a left attachment path joining
the two left attachment points and in some subgraph , . It is useful here to refer
again to Figure 9, where we now replace by , and we imagine the left
attachment path as joining
to the point (where neither nor is visible in the Figure). Since
is on this , we now reroute (with the help of Lemma 4.9a) so that arrives at and from there reaches along the path incident on . The arrival at is accomplished by viewing
as the left attachment
point of the square containing it, that is by observing that , and
then applying Lemma 4.9a. Explicitly then, we replace
the subpath
of the current by the path , leaving the rest
of unchanged. Call the new cycle
.
We now also reroute the path so that it now intersects at , calling the rerouted
path , as follows. Now
will still begin at
, with the goal of intersecting at . We do this by moving through a sequence of right
attachment paths, using Lemma 4.9b, to
reach (similarly
regarded as
in applying Lemma 4.9b). Then reaches along the path incident to . Explicitly we replace the
subpath
of by the path .
We leave path the same,
and similarly now modify so
that it reaches the point on some attachment path , to obtain the new cycle . We also make an adjustment
to similar to the one made to
so that the new intersects at , using Lemma 4.9b again.
The final cycle and
paths , , are as follows, with remaining the same. To distinguish
the two copies of in the
discussion, the one contained in will have its attachment points and
paths suffixed by a , while the
one contained in will have
these points and paths without a suffix.
We thus obtain a new , consisting of the final and the three independent
paths , , . Finally we note that very
similar adjustments can be made if is on some right attachment
path in some subgraph , . We omit here the proof of
independence of the new paths. Henceforth we will assume in all our
cases that and belong to squares. The
arguments for these points belonging to attachment paths are similar to
the ones given above, obtained by simple changes to the ones involving
squares.
Still with , consider the
possibility .
We consider the two cases where first is a point in some copy of a
square , and second
where is not such a point
and is therefore on some attachment path in some copy of .
So assume is in some copy
of and by symmetry we can
take . This case is
illustrated in Figure 10.
Figure 10
Since corresponds to in , then belongs to a corresponding copy of . We then form a cycle
(shown in bold
in Figure 10) as follows,
Observing that does not enter
, we can take , and obtain the following
three independent paths
(illustrated in dashed curves), where again paths numbered , , in the figure correspond to paths in the description which
follows.
Still under the case suppose that is on some attachment path in some copy of , as illustrated in
Figure 11. The reader can consult this
figure as they read the analysis of this case.
Figure 11
Again we may suppose by symmetry that and that , where is the corresponding copy of in . Here lies on the attachment path
corresponding to in , where the tuples and over differ only in their final
letter, that being in and in . We have labeled the two subgraphs
and of in that figure. Note that joins these two subgraphs at
their left attachment points.
Now in we can form
the cycle and three independent
paths as follows, where
know that since
never enters .
This completes consideration of the case . The case is argued similarly, requiring
several subcases. The missing edge is now no longer on a path in . The cases involved still use the
the decomposition of into
the subgraphs
together with the various attachment paths and the connection paths in
and . In each case, we find an in by finding a cycle , a point and three independent paths. We omit the arguments here,
and instead provide the details in Appendix II.
Corollary 4.11.
For even and , .
Proof. Since is
-saturated by Theorems 4.3 and 4.10 it suffices to show
that . We continue with the notation and from the preceding theorem.
For any connected graph let
. We say
that is an
excess set of if is a spanning tree of
, so that for such an we have . By Lemma 4.1 we know that
is a forest, each component
tree of which is pendant on .
So any edge of is a cut edge
of . Therefore for any excess
set of we have . So in
fact must be an excess
set of . So we let , and we
determine from which we
obtain , since is a
connected spanning subgraph of .
Consider first . Now , and contains a unique cycle (the
outside cycle). Take the set , where (resp. ) is an edge on the unique cycle of
(resp. ), and is any edge of , say of . Then we can see that is a spanning tree
of as follows. First note
that spans
. Since and are connected and unicyclic,
then and are spanning trees of
and respectively. Also is a path joining these two
spanning trees, thus yielding that is a spanning tree of . So is an excess set for , and hence .
We claim that the sequence satisfies the recurrence . For this, it
suffices to find an excess set of of size .
By construction we have . Let
(resp. ) be an excess set in
(resp. ) of size , and let be any edge in . Letting , we see that is a spanning tree
of as follows. This is
because whichever of or
that does not contain is then a path in joining the two
spanning trees and of and
respectively. So is an
excess set of , and has
size , as
required.
Finally a simple induction using this recurrence and shows that is the unique
solution to the recurrence for . This yields as required.
5. Conclusions
In this paper we studied the saturation function ; that is, the
minimum number of edges in any saturated spanning subgraph of
the -dimensional grid . Here is the graph family of all
subdivisions of the complete graph . In dimension 2 we obtained the
exact result if at least one of or is odd with and . In dimension 3 we found the near
exact result for even with . For arbitrary with even and , we proved that .
The upper bounds for dimension were obtained by the inductive construction of the -saturated graph described in
the paper.
6. Appendix I – independence of
paths
In this Appendix we prove independence of the three paths used in the
construction of an in the
proof of Theorem 4.10, in the case and .
First we show that . The initial subpath of satisfies . The remaining succession of paths on , which we call , satisfies
lies in . So it remains to show
that . The subpath of
is a right attachment path. But consists of a left attachment path (so it has empty
intersection with ), a path
in avoiding (so having empty
intersection with ), and
finally a connection path which must have empty
intersection with since
contains no connection
points and does not contain . We see that the subpath of lies in . But then Lemma 4.9c,d give that . So we get as desired.
Next we show that . Since , the initial subpath of has empty
intersection with . The second
subpath of intersects at the vertex by construction. Hence as
required.
Now we show that . Here the proof is similar to the one given
for showing . First observe that the initial subpath has empty intersection with by construction. The rest of lies in , so it suffices to show that .
By construction . Suppose first that . Then
the subpath. of has empty intersection with
, since is a left attachment path while
is a right
attachment path followed by a path not containing . It then remains to
observe that the subpath intersects
the subpath at by Lemma 4.9c,d,
showing that as desired. Now suppose that . The we
have . So
the path on reaches , and
again we get .
To complete the proof of independence of the three paths, we show
that , where the containment is
clear by definition. We easily have , since , while . As for , since and since neither nor contain connection points of , it suffices to show that
and . The first claim is clear since by definition . For the
second claim, we note that . But is a right attachment path, which therefore cannot contain
, followed by a path
internal to which
avoids by Lemma
4.9d. So we obtain , and
it follows that , yielding the independence of the three paths , .
7. Appendix
II – completion of proof of theorem 4.10
In this Appendix we complete the proof of Theorem 4.10 by providing the
details for the case , where
the nonedge being added to has dimension in . The case in the proof of Theorem 4.10 was treated in the
main body of the paper.
First we can reduce to the case where . Suppose not. If exactly one of or is in , then is an -dimensional nonedge following or
preceding some path ,
a contradiction to . If
and , then and , showing that and disagree (in fact by more than ) in the ’st coordinate, so contradicting
. Finally with either or , then induction
immediately gives an in
either or and therefore in , as required.
From it follows from the decomposition of
at the beginning of the
proof that . We cannot have since
neither nor contain nonedges of , while points of differ from points of in more than one coordinate.
Recalling the notation we must have either
both and in , or
one of in while the other is in
.
We begin by assuming , and
later consider the case .
Suppose first that as in case , and we treat the case afterwards. Since , then and are on distinct -paths in dimension , say and , since otherwise and would disagree in coordinate , contrary to . By possibly renaming we can
suppose . Now
(resp. ) disagrees with (resp. ) only in the ’st coordinate, but and disagree (by ) only in their ’th coordinate. Therefore and disagree by only in the ’th coordinate. Then there are two
possibilities under case .
and are successive points on some
-path of dimension , or
and are successive points on the copy
of or the copy of in .
(resp. ) of (resp. ) in .
Consider first possibility case , shown in Figure 12. Then we can take
with . So we must have , and by symmetry we
can take . So we have that
and are in the same component of as , so .
Figure 12 An in where , successive points on a
-path of dimension
We claim that one of is the head of its -path in dimension , while the other must be the tail of
its -path of this dimension, For
this, note that and are neighbors in so are on opposite sides of
the bipartition of
in step of the construction of
. So by step of that construction, and receive opposite head/tail
designations for their -paths
dimension , proving the
claim.
So arbitrarily take , forcing . Therefore is
in the same tree component of as , so we get since .
Now let and , so . Then
and are successive points on a
path , for
some , corresponding
to . (We note here
that in possibility (2), and are successive points on
either or , which will be considered
when is analyzed.) Since and , together with their heads
and tails, are corresponding -paths of dimension in and respectively, we have by step of the construction of . So the point corresponding to is , noting also that . Hence , and by
renaming we can let . So and still recall that . The various -paths and points are illustrated in
Figure 12.
We now form the cycle in
as follows.
We then take , and form the three independent paths in as follows.
Consider now possibility case , still in case and with . Here and are successive points on or on , say by symmetry on . So and are corresponding successive
points on . As before we
can take and . This case is illustrated in Figure 13.
Figure 13
First we form the cycle , indicated in bold in that
figure, as follows.
Now we take , and form three independent paths as follows.
Still assuming , consider
case ; that is, where one of
is in while the other is in
and illustrated in
Figure 14. By symmetry we can suppose
, the latter having
endpoints and , the case being essentially the same.
Now where , and by symmetry we
take .
Figure 14
We can further specify as
follows. Now implies and . Since and , we then obtain . So this
forces to be the point on
adjacent to .
Since , then
because is even. Hence , so by construction
. Then
in we have have a cycle
, the point , and three independent
paths , , as follows. In Figure 14,
is the -cycle illustrated in bold lines, and
the paths in dashed lines.
This completes the analysis of the possibility , and we now move to the assumption
. Again we begin with case
, where . Figure 15 illustrates this case. The
labeled vertices in this figure are explained in the discussion which
follows, and the reader may wish to follow this discussion using the
figure for illustration.
Figure 15 , and on distinct -paths of dimension
Let
correspond to and in ; that is, while
and for . Similarly let be such
that while and for .
Since , we see that and are on distinct -paths of dimension in , while and are on the corresponding distinct
-paths of dimension in . Now and are neighbors in , so are on opposite sides of
the bipartition of
. So and receive opposite head/tail
designations in and
respectively. By
symmetry we can take so , so and . Therefore and and and are neighbors in with and .
Now let
be the vertex with and for . Similarly let (resp. ) be such that and
for (resp. and for
). So we have , and
by symmetry we can take . So we have .
Again, the situation is illustrated in Figure 15.
We claim that .
For this it suffices to show that since and are joined by a path (in fact he
subpath subpath of
) in . Since we have
. So since and are on opposite sides of the
bipartition , it
follows that and
are also on
opposite sides of the same bipartition. So we get since . But since by construction, we get . It
follows that , so finally . Again see Figure 15.
We can then form our cycle
(shown in bold in Figure 15 ) and
point , together
with three independent
paths (shown dashed in that figure) as follows.
Still with , we finally
consider case , where one of
is in while the other is in
. One of the subcases
here described later is illustrated in Figure 16. That figure
is still useful for the reductions which follow.
By symmetry we can take , recalling that the path has ends and . Also where we can take
or , and we can arbitrarily
take . With the same
argument as in the case , we
find that must be a neighbor
in of across an edge of dimension . A straightforward induction on shows that for . Hence is a neighbor of in . Observe further that every
neighbor of in across an edge of dimension lies on some left attachment path
for some length string over , while for dimension the two neighbors of are in the square containing .
We are therefore reduced to the following two cases defined by :
is the point adjacent
to on some attachment path
, or
belongs to no
attachment path, in which case is one of the two neighbors of
in .
Consider first case a), illustrated in Figure 16 with details
which follow. Then is
in a subgraph . In Figure 16 we illustrate the subgraphs and by two smaller rectangles inside the
lower left larger rectangle representing . The ends of the attachment path
are shown as and , while the ends of as and . (Though we don’t use this in the construction which follows,
we can deduce by observing that
every vertex in has the same
length suffix as , so . From this and we can deduce the coordinates , and .)
We now have in the
cycle , the point , and three independent
paths as follows and illustrated in the
Figure 16.
Figure 16 case where and
Now consider the second case b); that is, where belongs to no attachment path and
is adjacent to in . Observe here that . This is
because , so (since is even),
so since
is a neighbor of . Now in we form a cycle , the point , and three
independent
paths as follows.
N. Alon. An extremal problem for sets with applications to graph theory. Journal of Combinatorial Theory, Series A, 40(1):82-89, 1985. https://doi.org/10.1016/0097-3165(85)90048-2.
L. B. Bollobas. On a conjecture of Erdos, Hajnal, and Moon. American Mathematical Monthly, 178-179, 1967.
J. Balogh, B. Bollobas, H. Duminil-Copin, and R. Morris. The sharp threshold for bootstrap percolation in all dimensions. Transactions of the American Mathematical Society, 364(5):2667-2701, 2012. https://doi.org/10.1090/S0002-9947-2011-05552-2.
C. A. Barefoot, L. H. Clark, A. Depew, R. C. Ertinger, and L. A. Szekely. Subdivision thresholds for two classes of graphs. Discrete Mathematics, 125(1-3):15-30, 1994. https://doi.org/10.1016/0012-365X(94)90140-6.
B. Bollobas. Determination of extremal graphs by using weights. Wiss. Z. Techn. Hochsch. Ilmenau, 13:419-421, 1967.
B. Bollobas. Weakly k-saturated graphs. In Beitrage zur Graphentheorie (Kolloquium, Manebach, 1967), volume 25, page 31. Teubner, Leipzig, 1968.
B. Bollobas, M. Przykucki, O. Riordan, and J. Sahasrabudhe. On the maximum running time in graph bootstrap percolation. arXiv preprint arXiv:1510.07096, 2015. https://doi.org/10.48550/arXiv.1510.07096.
J. Chalupa, P. L. Leath, and G. R. Reich. Bootstrap percolation on a Bethe lattice. Journal of Physics C: Solid State Physics, 12(1):L31, 1979. https://doi.org/10.1088/0022-3719/12/1/008.
G. Chartrand, L. Lesniak, and P. Zhang. Graphs & Digraphs. Chapman Hall/CRC, 6th edition, 2015. https://doi.org/10.1201/b19731.
S.-y. Choi and P. Guan. Minimum critical squarefree subgraph of a hypercube. In Proceedings of the Thirty-Ninth Southeastern International Conference on Combinatorics, Graph Theory and Computing, volume 189, pages 57–64, 2008.
R. Diestel. Graph Theory. Springer-Verlag, New York, 1st edition, 1997.
P. Erdos. On some problems in graph theory, combinatorial analysis and combinatorial number theory. Graph Theory and Combinatorics (Cambridge, 1983):1–17, 1984.
P. Erdos, A. Hajnal, and J. W. Moon. A problem in graph theory. The American Mathematical Monthly, 71(10):1107–1110, 1964.
J. R. Faudree, R. J. Faudree, and J. R. Schmitt. A survey of minimum saturated graphs. The Electronic Journal of Combinatorics, 1000:DS19–Jul, 2011. https://doi.org/10.37236/41.
M. Ferrara, M. Jacobson, K. G. Milans, C. Tennenhouse, and P. S. Wenger. Saturation numbers for families of graph subdivisions. Journal of Graph Theory, 71(4):416-434, 2012. https://doi.org/10.1002/jgt.21625.
M. Ferrara, M. Jacobson, F. Pfender, and P. Wenger. Graph saturation in multipartite graphs. Journal of Combinatorics, 7, Aug. 2014. https://doi.org/10.4310/JOC.2016.v7.n1.a1.
R. J. Gould. Developments on saturated graphs. Chapman and Hall/CRC, 2019, pages 111-133.
A. E. Holroyd. Sharp metastability threshold for two-dimensional bootstrap percolation. Probability Theory and Related Fields, 125:195-224, 2003. https://doi.org/10.1007/s00440-002-0239-x.
O. Janzer. The extremal number of the subdivisions of the complete bipartite graph. SIAM Journal on Discrete Mathematics, 34(1):241-250, 2020. https://doi.org/10.1137/19M1269798.
T. Jiang and Y. Qiu. The extremal number of bipartite subdivisions. SIAM Journal on Discrete Mathematics, 34(1):556-570, 2020. https://doi.org/10.1137/19M1265442.
J. R. Johnson and T. Pinto. Saturated subgraphs of the hypercube. Combinatorics, Probability and Computing, 26(1):5-21, 2017. https://doi.org/10.1017/S0963548316000316.
L. Kászonyi and Z. Tuza. Saturated graphs with minimal number of edges. Journal of Graph Theory, 10(2):203–210, 1986. https://doi.org/10.1002/jgt.3190100209.
W. F. Klostermeyer and A. Yeo. Edge domination in grids. Journal of Combinatorial Mathematics and Combinatorial Computing, 95:99–117, 2015.
R. Lang, H. Lei, and J. Zhang. On the saturation spectrum of families of cycle subdivisions. Theoretical Computer Science, 962:113941, 2023. https://doi.org/10.1016/j.tcs.2023.113941.
Z. Miller and G. Yane. A saturation problem on meshes. Submitted.
N. Morrison, J. A. Noel, and A. Scott. Saturation in the hypercube and bootstrap percolation. Combinatorics, Probability and Computing, 26(1):78–98, 2017. https://doi.org/10.1017/S0963548316000122.
G. Moshkovitz and A. Shapira. Exact bounds for some hypergraph saturation problems. Journal of Combinatorial Theory, Series B, 111:242–248, 2015. https://doi.org/10.1016/j.jctb.2014.08.004.
O. Pikhurko. Extremal hypergraphs. University of Cambridge, 1999.
D. J. Watts. A simple model of global cascades on random networks. Proceedings of the National Academy of Sciences, 99(9):5766–5771, 2002. https://doi.org/10.1073/pnas.082090499.
W. Wessel. Über eine Klasse paarer Graphen. I. Beweis einer Vermutung von Erdos, Hajnal und Moon. Wiss. Z. Techn. Hochsch. Ilmenau, 12:253–256, 1966.
W. Wessel. Über eine Klasse paarer Graphen. II. Bestimmung der Minimalgraphen. Wiss. Z. Techn. Hochsch. Ilmenau, 13:423–426, 1967.
D. B. West. Introduction to Graph Theory, volume 2. Prentice Hall, Upper Saddle River, 2001.
A. A. Zykov. On some properties of linear complexes. Matematicheskii Sbornik, 66(2):163–188, 1949.