1. Introduction
In their seminal paper, Bruck and Bose [1] proved that every finite translation
plane can be obtained from a construction that starts with a
spread of an odd-dimensional projective space.
Letting that space be , a
spread is a partition of the space into pairwise disjoint -dimensional subspaces. Much of this
work was anticipated in 1954, by André [2] who provided a similar construction of finite
translation planes using a vector space model that was later shown to be
equivalent to the projective space model developed by Bruck and Bose.
Moreover, André provided a robust construction of spreads of for any integer and prime power which yields distinct, though possibly
isomorphic, spreads, which comprised one of the first known infinite
families of non-Desarguesian translation planes.
In addition to the André planes, the early days of the study of
finite projective planes yielded several infinite families of
translation planes, as well as a number of sporadic examples with
particularly interesting automorphisms. Primarily in the 1990s and into
the 2000s, there was an explosion in both construction techniques for
new planes, as well as in computational results classifying the finite
translation planes of orders 16 (Dempwolff and Reifart [3]), 25 (Czerwinski and
Oakden [4]), 27
(Dempwolff [5]) and
49 (Mathon and Royle [6]). Ironically, this mass of data seems
to have provided a disincentive to continued work. On the one hand, the
plethora of new construction techniques gives the impression that most
translation planes are “known”, while the computational results suggest
that a complete classification of finite projective planes, let alone
translation planes, is infeasible.
Indeed, the André planes provide an example of this problem in
microcosm. As described by Johnson, et al. [7], the André planes are all “known,” in the sense
that for any order it is
straightforward to construct all André planes; there are no new André
planes to discover. But with modern computing power, it is now possible
to construct and compute with models for all André planes of order up to
and including 125. Looking to the future, the next order for which André
planes exist is 169, and while it is possible to construct all André spreads from a regular
spread of currently, the
computational power required to determine isomorphism between more than
4000 projective planes of order 169, or even the equivalent problem of
sorting isomorphism between more than 4000 spreads of , is certainly formidable. Our
goal in this paper is to develop algorithms using group actions with a
greatly reduced degree to determine representatives for all isomorphism
classes of André planes, as well as more efficient algorithms that count
isomorphism classes of André planes non-constructively with Burnside’s
lemma.
We believe that the computational enumeration of existing families of
translation planes, to truly measure the robustness of known
construction techniques against the existing classification results,
represents a fruitful way forward in the theory of finite projective
planes. In addition to the censuses of translation planes cited above,
Moorhouse [8] has
created an online database which contains projective planes of order up
to 49, including the planes from existing censuses, as well as a large
number of planes obtainable from known planes using derivation and
dualization. Using this resource as a reference, the author [9] has created all of the
translation planes of order 25 from known construction techniques and
correlated them against the list of translation planes of order 25
determined by Czerwinski and Oakden [4]. This analysis showed that there is a
translation planes of order 25, designated by Czerwinski and Oakden, which is
not part of any known infinite family of translation planes. The author
was able to provide some theoretical context around how might arise, but it remains an open
question whether this plane is sporadic, or if it is part of an infinite
family yet to be found; regardless, it provides an interesting question
for further research.
2. Regular Spreads
In this section, we provide a framework for addressing the André
planes obtained from .
Many of the results here are well known for the case , and for the higher-dimensional cases
many of these results were certainly known to Bruck [10], but were never stated
and proved. Regardless, there is value in clearly articulating the
results here in a unified manner, so that we may refer to them as needed
in what follows.
Let be an integer and
a prime power. Define , , and let
denote the set of non-zero elements of , with defined analogously. Let be a -dimensional vector space over
. This vector space is a model for using homogeneous coordinates,
such that and define the same point for all
.
A spread of is a set of pairwise disjoint -dimensional subspaces which partition
the points of . A
regulus in
is a set of pairwise disjoint -dimensional subspaces such that any
line that meets three spaces in
meets all of them. Bruck and Bose [1] show that any three pairwise disjoint
-dimensional subspaces lie in a
unique regulus, and define a regular spread to be
one that contains the regulus generated by any three of its elements.
Regular spreads are significant in that for any , the Bruck and Bose construction
of translation planes yields a Desarguesian plane if and only if the
spread used to construct the plane is regular. From this definition, it
is easy to see that any two regular spreads that meet in at least three
subspaces must meet in at least an entire regulus; Bruck and Bose show
that a regulus and a single space disjoint from all elements of the
regulus uniquely determine a regular spread, which shows that spaces is the largest possible
intersection between two distinct regular spreads.
Bruck and Bose give a specific coordinate model for a regular spread
which will be very useful in what follows. Define , and for
any , let . These are all
-dimensional subspaces, and are
pairwise disjoint, hence we have a spread ; Bruck and Bose prove that this spread is regular.
Let denote the group of
field automorphisms of with fixed
field , which necessarily has
order , and let be any element of . It is clear by linearity of
that the sets
are also -dimensional subspaces of
. André [2] noted that for any , the set of subspaces ,
where is the norm function
from to , can be replaced by the set , for any to
create a new spread of
that is not regular. Moreover, each can be replaced (or not) independently as varies over , yielding different, though potentially
isomorphic, spreads. These spreads are called André
spreads, and the translation planes that arise from them are
called André planes. The sets , and any set of spaces in isomorphic to them, are called
norm surfaces in .
An important notion when dealing with André spreads is the concept of
a linear set of norm surfaces. Bruck has developed
an abstract definition of linearity, but ultimately a set of norm surfaces in is linear if there exists a
collineation of which
leaves invariant and maps
onto a set of norm surfaces . André spreads are exactly
those that are obtained by replacing norm surfaces from a linear set;
generically a spread obtained from replacing an arbitrary set of norm
surfaces in a regular spread is called subregular.
We will shortly recall some of Bruck’s results about linearity which
require the abstract definition to prove, but we only need the cited
results in what follows.
For the case, Bruck [11] extensively developed an
isomorphism between the regular spread of and the
Miquelian inversive plane . For
the higher-dimensional cases, Bruck [10] generalized this connection to
higher-dimensional circle geometries, keeping intact the isomorphism
between a regular spread of and . The key takeaway from the circle
geometry connection is that it shows an isomorphism between the regular
spread of and the projective line , coordinatized as , where the spaces in
map to the points of , and the reguli in map to order sublines in . Moreover, there is a
homomorphism from the group of
collineations of leaving
invariant onto ; this allows us to
provide an explicit description of the group of collineations of leaving invariant. Bruck [11] proved the following result
for the case, and Bruck [10] stated that the
corresponding result is true for all ; a detailed proof was provided by the
author [12].
Proposition 1. Let be a regular spread in . Then the group of
collineations of that
leaves invariant consists
of the transformations acting via
Moreover, the subgroup of this group that leaves each element of the
spread invariant is exactly
the set of transformations
Reguli and norm surfaces are the same when , but in higher dimensions, the
isomorphism between the regular spread of and maps norm surfaces onto
structures in that
Bruck calls covers. Bruck [10] shows that every cover satisfies an
equation of the form or , where , and . For fixed , or
for fixed , the set of covers
obtained by varying over is a linear set
of covers, which corresponds to a linear set of norm surfaces of . The two points not covered by
these sets ( in the first
case, in the second) are called
the carriers of the linear set.
Bruck has proven the following result about linear sets, which are
collected here to illustrate an important difference between the and higher-dimensional cases:
Proposition 2. Let be a norm surface in a regular
spread of with . Then has a unique pair of carriers,
and lies in a unique linear set of norm surfaces. If , then the lines of off are partitioned into pairs of
carriers of , and any norm
surface (regulus) disjoint from lies together with in a unique linear set.
With this result in place, we can prove an important result about how
two norm surfaces/covers can intersect; the case is an obvious corollary of the
fact that norm surfaces are reguli, while the author [12] proved the case. This result feels like it
should be known, but may have fallen through the cracks based on its
trivial reduction in the
case.
Proposition 3. Let and be distinct norm surfaces in a
regular spread of , each consisting of elements of the
spread. Then and can intersect in at most elements of the
spread.
Proof. For ease of notation, let and let . As discussed
above, determining the number of spread elements in which two norm
surfaces in the regular spread intersect is equivalent to determining the size of the
intersection of two covers and
in . Since all covers are
isomorphic, we may assume that
has equation . We
assume has equation for some
, ; the argument for the case
where has equation is nearly identical and
slightly easier. Suppose . We know and
since , and since . Note that this latter equation
has exactly solutions since
is a cover, and thus is not
identically zero.
Expand the second equation to yield:
Now multiply on both sides by , which is not identically zero, to
obtain:
Since , we can simplify
to obtain:
This equation yields a polynomial expression in of degree at most satisfied by all elements of , and is not identically
zero. Thus there are at most
values for satisfying this
equation, proving the result. 
A second key difference between the and higher-dimensional cases is the
number of potential replacement sets for a norm surface. When sorting
isomorphism classes, it becomes important to understand what is
happening with all potential replacements under collineations of . To this end, recall that for
any we have
is an -dimensional
subspace, and our replacements for are the sets . Define , and let be the collineation
defined via . Clearly leaves invariant and maps onto , hence each of the is a regular spread, and
there is a collineation of that maps to that has the net effect of
replacing all of the spaces of the norm surfaces in with .
Our final general result describes the stabilizer groups associated
with various collections of the norm surfaces in .
Theorem 1. In , where and is a prime power, and , let be
the regular spread with linear set . Then
The collineation group of leaving the set of norm surfaces invariant
consists of the collineations ;
the permutation action of acting on the sets is action-isomorphic to the group of transformations acting via ;
the collineation group of leaving the set of norm surfaces invariant
consists of the collineations ; and
the permutation action of acting on the sets is
action-isomorphic to the group of transformations acting via .
Proof. Suppose is
a collineation of that
maps onto itself. Since
maps spaces of onto spaces of , must leave invariant. So by Proposition 1
for some
, and . Moreover, since leaves invariant and invariant, it must leave invariant, meaning
must either leave both and invariant, or interchange them. In
the former case, we must have , while in the latter we must have
, hence must be in the set of collineations
.
Let . If ,
then . Note that for some , and there exists such that for all . So
if , then we have
.
As and depend only on , , and , for all we must have showing that
leaves invariant. The
calculation for is similar, and these results together show
that is the entire group
of collineations of
leaving invariant.
Moreover, these calculations show half of the claimed action
isomorphism, namely that every element of acts on as some . To show that all
such elements can be obtained, note that acts as for any with , and acts as for any with .
Let be the group of
collineations of that
leave invariant, and
let . Suppose first that
. In this case, since , contains at least three
reguli lying in and at
least three reguli lying in . Thus must map
at least two reguli and of onto reguli of either or . Since
and are disjoint reguli, they contain
lines, forcing to map onto either itself or onto ; the analogous statement for
implies leaves the set of regular spreads
invariant. If
, the argument is similar,
but slightly easier and works for . Since any norm surface has more than spaces it lies in a unique regular
spread, hence if maps to , it must map to ; thus must leave the set of regular
spreads invariant.
Since the group of all collineations of leaving invariant has order , the stabilizer of in is certainly no bigger. We also know
that the orbit of under
has size at most , hence is at most . Now, consider ,
where and , and first look at
the case where for and .
Since is non-zero, we can can
write for some
. Looking at an
arbitrary point of , we
have This shows that maps every space to . For all
with fixed norm , we
have ; moreover, since is an automorphism of , there exists an automorphism such that for all . Thus we find maps to for some , , and
thus leaves
invariant.
Now we address the case where for some and .
Here is non-zero, so we can write
for
some . Again we
calculate Using calculations similar to the above, and
again letting such
that for all , we see that maps to , and thus also leaves invariant.
We have shown that . To show equality, note that ostensibly has elements, but there
could in principle be some collapsing wherein . But if this occurs, we have . Both and leaves invariant, which forces to leave
invariant as well, but this
only happens if , in which case . Thus , proving it is the entirety of .
The above calculations substantially show the action isomorphism of
part 4 as well, since it shows that every acts on in the fashion of for
some , and . As above, we note
that acts as for any
with , and acts
as
for any with . Hence there is an
action isomorphism between and ,
completing the proof. 
3. Two-Dimensional André Planes
Now that we have our basic machinery in place, we begin our
enumeration with the case,
two-dimensional André planes. Refreshing terminology, let be a regular spread of , a prime power, and let be a linear set of reguli in . Letting vary over all subsets of
, we can create every
two-dimensional André plane of order with the set of spreads
We call the size of the set the index of the André spread. An
André plane of index either 0 or is obviously regular. Albert [13] has shown that any plane
obtained by switching a single regulus in a regular spread is
necessarily the Hall plane, thus any André plane of index either 1 or
is a Hall plane. In what
follows, we exclude these cases from consideration.
Since we are interested in sorting isomorphism classes, we may
further restrict our attention to André spreads of index at most . Under the collineation
, is isomorphic to
so every equivalence class of isomorphic André planes
contains at least one representative of index at most .
These constraints show that the only André spreads of and are the regular spread and the
Hall spread. Thus we may restrict out attention to André spreads of
with , and of index .
The next two results are highly reminiscent of the work of
Walker [14],[15]. In those papers,
Walker shows that the group of automorphisms that leaves a subregular
spread derived from a regular spread invariant almost always leaves the
original regular spread invariant as well, with the exceptions being
André spreads with index either 1 or . Our problem is closely
related, but not identical, and we need to go into the details of the
index case in order
to sort isomorphisms, so we prove the needed results directly. We begin
with a straightforward lemma which describes how a regular spread can
meet an André spread.
Lemma 1. Let be an André spread, with index satisfying in , . The regular spread meets in lines,
meets in lines, and no other
regular spread meets in more than
lines.
Proof. The intersection sizes of with and are simple consequences of the
construction of , and the lower bounds on those sizes are a simple
consequence of the bounds on . Let
be a regular spread
distinct from and . We can write as the union of two
partial spreads . Two distinct regular spreads can meet in at most lines, so meets each of the two components
of this union in at most lines,
making the total size of the intersection at most . 
With this lemma in place, we can prove our key result to sort
isomorphisms between two-dimensional André planes.
Proposition 4. Let and be André spreads of
derived from the same
regular spread , with
indices between and , inclusive. Then and are isomorphic if
and only if:
there is a collineation of leaving invariant that maps the set of
reguli onto the set of reguli ; or
there is a collineation of leaving invariant that maps the set of
reguli onto the set of reguli
Proof. The reverse direction is relatively straightforward:
clearly if there is an automorphism of that maps the set of reguli onto the
set of reguli , it is an explicit isomorphism from to . If there is an
automorphism of that maps to , then
maps to , since interchanges and for all , resulting in a spread derived
from with exactly the
reguli in reversed, namely .
Now suppose
and are
isomorphic, with collineation
mapping to
. By Lemma 1,
each of and
meets and in more than lines, and meet no other regular
spreads in that many lines. Hence must either leave and invariant, or must interchange
them. The former case forces and to have the same index, while the latter forces
and to have indices
summing to . By hypothesis the
indices of and
are both at
most , hence this case
only occurs if
and both have
index .
Suppose first the isomorphism leaves invariant. Then maps to , and
thus must map the set of reguli in onto a set of reguli contained in the
union of lines of the reguli in . Since has at
most reguli, by the
pigeonhole principle, the image of each regulus in under
must meet some regulus in
in
at least 3 lines, forcing it to be identical to one of those reguli.
Hence every regulus in must map under onto a regulus in , and the
fact that the indices of the two spreads are the same forces to map the set of reguli onto the
set of reguli .
If the isomorphism
interchanges and , we know that the index of
and is . Consider the collineation
. For any
, is a
regulus in and
also in . Applying to this regulus gives a regulus
, which lies in both and . Thus must be
a subset of the union of the reguli . But since the index of
is , this union is of reguli, and as before must share at least 3 lines with one of
the for , and
thus must be equal to that regulus. Therefore, is a collineation of leaving invariant that maps the set of
reguli onto the set of reguli , completing the
proof. 
Suppose and are two sets of reguli with size between 2 and
inclusive, in a
linear set of a regular
spread , and we wish to
determine if they generate isomorphic André spreads, and thus isomorphic
André planes. From Bruck [11]
is the only linear set of reguli containing , and
also the only linear set containing . Moreover, since and have less than elements, the complementary
sets of reguli and also have at least two
elements, and thus are only contained in the linear set . Thus any collineation of leaving invariant which maps to or its
complement in must leave
invariant. This allows us
to apply Theorem 1, parts 1 and 2, from which we find that
the spreads
and are
isomorphic if and only if there exist and such that , for some choice of
sign, or in the case where , there exist and such that , for some choice of sign.
For , this problem is
tractable by hand, since the only case we have to deal with is that of
index . One
possible set of size 2 is . There are no nontrivial automorphisms of , so we see this set maps to , and under for . The
inversion map
interchanges with and with , and the complement of each of these sets is already
represented, so these four sets form one orbit under . It is easy to see that the remaining
two pairs,
and , form a
second orbit. Hence there are two André planes of order 25 with index 2.
This is validated by the enumeration of all translation planes of order
25 by Czerwinski and Oakden [4], where 5 subregular spreads are found in
: the regular spread, a Hall
spread, a subregular spread from a non-linear triple, and two André
planes, one of which is in fact the regular nearfield plane of order 25.
Using MAGMA [16], we have
automated this calculation for small ; the code implementing this enumeration
is in Appendix A. For some small values of , all of the two-dimensional André
planes, excepting the Desarguesian and Hall planes, can be obtained from
the sets of s in Table 1.
Table 1: Enumeration of André Planes of Order With Index at Least 2
5 |
2 |
|
|
7 |
6 |
|
8 |
3 |
|
() |
9 |
12 |
|
() |
|
|
|
|
|
|
|
|
|
|
11 |
42 |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
This method produces representative sets of reguli for each André
plane, but bogs down as
increases, due to the need to create the actual subsets of reguli for
each index; for example, enumeration with runs in less than a minute, but
increasing to starts increase
run-time significantly. If we are only interested in the number of
distinct André planes with a given index, we can appeal to Burnside’s
lemma to count the number of orbits under the group , with the exception of the André
planes of index ,
where complementation is again a confounding factor. Let us consider
this case in more detail.
Let be the set of all subsets
of of size . Each element of induces an action on through its action on the individual
elements of . Though it does not act element-wise, complementation is
also a permutation on , which we
denote as , acting on via for all . Hence we can perform the same
Burnside counting in this case, but we have to use the group generated by the elements of and . This turns out not to be as
taxing as one might fear, based on the following result.
Proposition 5. Let , a prime
power, and let be the group of
transformations acting on , with and . Consider the induced
group action of on the set
of subsets of of size , and let be the group generated by the
permutations in and , acting on . Then:
For all ,
and commute;
= ;
fixes some
if and only if all orbits
of , when acting on , have even length; and
if all orbits of ,
when acting on , have even
length, then fixes elements of , where is the number of orbits of when acting on .
Proof. The first two statements are clear, since the action
of on preserves complementation, namely if
then , and
has order 2 as a
permutation. Suppose now that for some we have . For each element , we have
in the action of on . Since ,
this implies for all , the preimage of under must be in , hence for all , we have . So if is any orbit of in its action on , its elements must alternate being in
and not in , forcing each such orbit to have even
length. Conversely, suppose every orbit of in its action on has even length. Pick one element
from each orbit of , and let
be the union of the orbits of
these elements under . This
set contains exactly half of the elements of , and maps each element of to an element not in , hence fixes . This also
shows the fourth claim, since each orbit of in its action on splits into two parts, each of which
can be picked independently of all other orbits, to add into a set fixed by . 
With this proposition in place, we have developed code in MAGMA to
implement this Burnside counting; this code can be seen in Appendix B.
For some small values of , we
obtain the counts of André planes with a given index in Table 2, but note that the algorithm scales
much better than the enumeration algorithm; for example, a run with
counts all non-isomorphic
André planes in less than a minute.
Table 2: Numbers of André Planes of Order , by Index
|
2 |
3 |
4 |
5 |
6 |
7 |
8 |
9 |
10 |
11 |
12 |
13 |
5 |
2 |
|
|
|
|
|
|
|
|
|
|
|
7 |
3 |
3 |
|
|
|
|
|
|
|
|
|
|
8 |
1 |
2 |
|
|
|
|
|
|
|
|
|
|
9 |
3 |
4 |
5 |
|
|
|
|
|
|
|
|
|
11 |
5 |
8 |
16 |
13 |
|
|
|
|
|
|
|
|
13 |
6 |
12 |
29 |
38 |
35 |
|
|
|
|
|
|
|
16 |
3 |
7 |
18 |
34 |
54 |
66 |
|
|
|
|
|
|
17 |
8 |
21 |
72 |
147 |
280 |
375 |
257 |
|
|
|
|
|
19 |
9 |
27 |
104 |
252 |
561 |
912 |
1282 |
765 |
|
|
|
|
23 |
11 |
40 |
195 |
621 |
1782 |
3936 |
7440 |
11410 |
14938 |
8359 |
|
|
25 |
8 |
30 |
143 |
487 |
1517 |
3741 |
7934 |
13897 |
20876 |
26390 |
14632 |
|
27 |
5 |
20 |
112 |
434 |
1532 |
4264 |
10145 |
20121 |
34291 |
49668 |
62220 |
33798 |
4. Higher-Dimensional André
Planes
The most significant difference between the two-dimensional André
planes and the higher-dimensional case is the presence of multiple
replacements for a norm surface, which in turn yields additional regular
spreads which can have a substantial intersection with an André spread.
This also makes the concept of index ambiguous, because just knowing
that a norm surface is reversed is not enough information to determine
an André spread; we need to know which replacement is chosen. To this
end, we define an indicator function ,
from which we obtain the André spread . Note that there are exactly such indicator functions, and
they define all of the possible André spreads.
Our first lemma is an analog of Lemma 1 for the
higher-dimensional case.
Lemma 2. Let be an André spread of , , , defined
via indicator function . Then the only regular spreads that meet in at least spaces are the regular spreads for in the range of . Moreover, a spread meets in either exactly
two spaces, or more than .
Proof. Let be as in the lemma statement, and let be a regular spread distinct from
the . Then for each
, meets the set in at
most spaces, hence the total
number of spaces in is at most . For any field
automorphism , if
for some
, then contains at least spaces
of ; otherwise
meets in just and . 
We can now prove the key result we need to sort isomorphism classes
of higher-dimensional André planes. Unlike Proposition 4, this result has
the advantage of providing a unified treatment of spreads independent of
the index; there is no special case for complementation since we are
keeping track of the replacement set for all norm surfaces in the linear
set. The downside is that this result is not as strong as Proposition 4,
requiring more group-theoretic computation during enumeration. However,
it is a distinct improvement over having to work with groups acting on
spreads of directly.
Proposition 6. Let and be non-regular André
spreads of , , , obtained from the regular spread with indicator functions and . Then any isomorphism between and must leave invariant. Moreover,
maps the collection of sets of
subspaces onto
itself.
Proof. Suppose and are isomorphic, with isomorphism a collineation in mapping onto . Since is an isomorphism, it must map the
set of regular spreads meeting in more than points
to the set of regular spreads meeting in more than points, hence must map the intersection of the
onto the
intersection of the . By Lemma 2, and are both subsets of
and since and
are not
regular, and
each contain
at least two spreads, implying the intersection is exactly , which must be left
invariant by .
Suppose there exists such that ; without loss of generality,
we may assume is the
identity. Then is a
collineation of that
leaves invariant, leaves
the set
invariant, and thus the linear set invariant as well. By Theorem 1, this implies and thus , showing leaves the set invariant.
Suppose now there is no such that
leaves invariant.
Since the set
of regular spreads meeting in more than points
has at least two members, there must be some spread in that meets in at least one, but
at most norm
surfaces, plus .
Without loss of generality, we may assume this spread is . Since , we know
and thus for some .
For each norm surface , is contained in the union of
at most norm surfaces
; cannot intersect as these two spaces
are left invariant by . But by
Lemma 2, must be identical with one of the , for otherwise
could only contain at
most spaces. By Proposition 2,
a norm surface belongs to only one linear set of norm surfaces in a
regular spread, so this implies that must map the set of norm surfaces
in onto the set of norm surfaces
in
.
Recall the collineation of defined via . Clearly
leaves each of and invariant, and maps onto for all . Thus leaves invariant, and leaves the set
invariant. Using
the same trick as before with Theorem 1, this forces
with either or with , and a similar calculation to the
above shows that maps the
collection of sets of subspaces onto
itself. 
In light of this proposition, we can use the group of Theorem 1 to sort
isomorphism between André spreads. Letting and be indicator functions of
two André spreads and , this proposition shows that there exists an isomorphism
between and
if and only if
there exists
such that
for some . From a
computational perspective, we can represent the indicator functions as
sets of ordered pairs and extending the action of naturally to these ordered
pairs, we see that and are isomorphic if and only if there exists such that .
Table 3: Enumeration of Higher-Dimensional André Planes for Small . O is the
Order of the Plane, # Is the Number of
Planes
2 |
3 |
27 |
1 |
|
|
2 |
4 |
64 |
2 |
|
|
2 |
5 |
125 |
6 |
, |
|
|
|
|
, |
|
|
|
|
, |
3 |
3 |
81 |
2 |
|
|
3 |
4 |
256 |
3 |
|
|
|
|
|
|
|
As before, this process is tractable for small cases by hand; let us
consider the André planes of order 27, whence and . In this case, there are only 9 André
spreads. Starting with , which is just the
regular spread , we see this
set is preserved under both multiplication by and inversion, so the only other
spreads in its orbit are and , namely and . A non-Desarguesian André plane
is represented by the set . Applying twice shows that and represent isomorphic
spreads, and applying maps to . Thus all six of the
non-regular André spreads are isomorphic, and there is just one
non-Desarguesian André plane of order 27.
Representing our sets as a sequence of field automorphisms by
defining a consistent ordering of the elements of , we have implemented this sorting
algorithm in MAGMA. Note that can be generated by four
elements: ,
, , and , where is the characteristic of , and we use MAGMA’s group generation
algorithms to create the entire group. Table 3 provides indicator
functions for all non-Desarguesian André planes for some small orders
with given and .
For larger values of , we can
again count André plane non-constructively using Burnside’s lemma and
the group of Theorem 1. The situation in higher dimensions is
messier than that with , as we
need to break into several cases to count the number of André spreads
fixed by each element of .
The cause of the difficulty is the inversion map ; when this map preserves both the spreads
and , but for larger this map may interchange some of the
replacements for norm surfaces, making the analysis more intricate.
Let us start with the easy case, namely , and define via
.
If is an André
spread left invariant by ,
and , then from the definition of , we know
For any , let be the length of its orbit under
. This implies that
for some automorphism . Since leaves invariant and contains only one
norm surface associated with norm , we must have . We see that , so if and only if is divisible by the order of as a field automorphism.
Therefore fixes an
André spread
if and only if all of the orbit lengths of are divisible by the order
of . If this does occur, then
much as in the case we can
count the number of André spreads fixed by . For one member of each orbit under , we may select an
arbitrary in
ways; then the union of the
orbits of under
, for each , form an André spread fixed by . Hence fixes André spreads
, where is the number of orbits
of .
If , the situation is slightly more
complex. Defining as
above, we have and
Let . If the orbit
length of under is even, then just like
before we can pick arbitrarily, and the orbit of under will form part of an André
spread. But if the orbit length of under is odd, we must have , or , which is eminently
feasible since is cyclic
of order . If is even, then for every there is a unique
such that , so any André spread
fixed by this must contain
for all in the orbit of under . If is odd, then for half of the , there is no such that , and those fix no André spreads. For the
remaining half there are two options for , giving two orbits of which can be in an André
spread fixed by .
Based on this analysis, we have developed code in MAGMA to implement
this counting procedure; this code appears as Appendix D. Table 4 shows
the number of non-isomorphic André planes, not including the
Desarguesian plane, for some small values of and . Note that one cannot say for a given
and that every André plane derived from an
André spread of has
dimension over a full kernel
; indeed the regular spread
belies this notion. We do not attempt to sort isomorphism between André
planes for different here, except
to comment that of the planes specifically enumerated in Tables 1 and 3, the only
isomorphism that occurs is in order 81 between the plane represented by
in Table 1 and the plane represented by in Table 3. This was
determined by analyzing the quasifields derived from the André spreads,
and in fact, this plane is a regular nearfield plane of order 81.
Table 4: Number of Non-Desarguesian André Planes Obtained From André
Spreads of
|
2 |
3 |
4 |
5 |
6 |
7 |
8 |
9 |
10 |
3 |
1 |
2 |
2 |
3 |
3 |
4 |
4 |
5 |
5 |
4 |
2 |
3 |
4 |
6 |
7 |
9 |
11 |
13 |
15 |
5 |
6 |
15 |
23 |
40 |
57 |
86 |
114 |
157 |
200 |
7 |
31 |
112 |
300 |
729 |
1503 |
2902 |
5134 |
8651 |
13795 |
8 |
25 |
114 |
402 |
1160 |
2877 |
6350 |
12804 |
24012 |
42445 |
5. Conclusion
Almost 30 years ago, Czerwinski and Oakden [4] used a computer search to determine the
complete list of twenty-one translation planes of order 25. Given the
plethora of construction techniques for translation planes now known, it
is a reasonable assumption that each of these planes belongs to a known
infinite family, but this turns out not to be the case. Upon initial
investigation by the author [9], the plane denoted by Czerwinski and Oakden required an
extension of a result of Baker, et al. [17] to explicate, and their plane still has not been placed in an
infinite family as of this writing. More than 25 years ago, Mathon and
Royle [6]
determined that there are 1347 translation planes of order 49. Almost
certainly there are translation planes discovered in this search that
can be analyzed to create new infinite families. And if not, McKay and
Royle [18] have
determined that there are 2833 two-dimensional translation planes (plus
the Desarguesian plane) of order 64. A researcher wishing to discover
new infinite families of translation planes has data in abundance to
analyze.
Yet our researcher finds themselves in an interesting dilemma, the
“unknown known.” There are translation planes that have been discovered
through computer search, but we have no idea if they belong to a known
infinite family. While there are some characterization results for
certain families of planes, there exists no complete dichotomous key
which allows one to start with a plane given as an incidence structure
and determine its provenance.
We believe that the best way forward to bridge the gap between our
theoretical canon and the paydirt that computer searches have provided
is to actually develop computational models of planes, family by family,
and correlate them against the search data we have. This paper provides
one step in this direction. Based on the results here, we have been able
to identify the translation planes of order 49 and 64 (two-dimensional)
that are André planes, with data given in Table 5, where the search
serial is the index in the appropriate reference of the plane in
question. The more we build out this table into a true database, the
easier it will be to identify those promising trailheads for the next
paths forward in the theory of translation planes.
Table 5: The Two-Dimensional André Planes of Orders 49 and 64
49 |
|
1344 |
7,375,872 |
897 |
|
49 |
|
1343 |
3,687,936 |
901 |
|
49 |
|
1345 |
3,687,936 |
899 |
|
49 |
|
1339 |
22,127,616 |
905 |
Regular Nearfield |
49 |
|
1324 |
7,375,872 |
917 |
|
49 |
|
1328 |
3,687,936 |
917 |
|
64 |
|
17 |
9,289,728 |
994 |
|
64 |
|
10 |
13,934,592 |
1042 |
|
64 |
|
16 |
9,289,728 |
1042 |
|