Voltage Operations
Voltage Operations
MAPS
ISABEL HUBARD
ELÍAS MOCHÁN
ANTONIO MONTERO
1. Introduction
original one, but with more ags. Thus, by applying one of these operations to
a polytope we get a new polytope with more ag orbits. Furthermore, it is not
dicult to see that, often, the local conguration of the ags does not depend
on the original polytope but on the operation per se. Could we use operations to
get any local conguration of the ag orbits? We shall explain this question more 5
precisely.
Operations on maps and polytopes have been studied previously. For example,
in [23], Orbani£, Pellicer and Weiss explore map operations to build k -orbit maps.
They take a combinatorial approach and divide each ag (incident triplets vertex-
edge-face) of a map into several new ones. In a very recent paper [6] Cunningham, 10
Pellicer and Williams introduce stratied operations to study monodromy groups
of some important families of non-regular maniplexes (a generalization of -the ag
graph of- polytopes, where some conditions are relaxed). This concept is closely
related to our concept of voltage operation and many of their results can be written
in our language and vice versa. However, not all classical operations have been 15
studied in this combinatorial form. As an example, we have the snub operation,
which is well known. The snub has the particularity that one should not take
all ags of the original polyhedron to obtain the operated one, but only half of
them. The previously studied operations assign one or more ags to each ag of
the operated object; voltage operations allow this to change, so we can obtain a 20
much broader class of polyhedra, maps, polytopes or maniplexes.
The ag graph of a polytope P is a properly edge-colored n-valent graph whose
vertices are the ags of P and two of them are adjacent if the corresponding ags
dier in exactly one element. The colors of the edges are given by the type of
element they dier in. The quotient of the ag graph of P by its symmetry group 25
is the symmetry type graph of P. The symmetry type graph can be thought as a
way to represent the local conguration of the ag orbits.
We can notice that when we apply a classical operation to dierent regular
polyhedra, the resulting polyhedra often have isomorphic symmetry type graphs.
(For example, ifMed represents the medial operation and P is a regular polyhedron, 30
the symmetry type graph of Med(P) will have either one single vertex, or two
vertices and an edge of color 2 joining them, depending on whether or not P is
self-dual.) Moreover, if we apply an operation O to two polyhedra, both with the
same symmetry type graph T , it is very likely that both resulting polyhedra have
the same symmetry type graph T .
′
35
A natural question to ask is: given a properly edge-colored n-valent graph T , is
there a maniplex whose symmetry type graph is T ? Going back to operations, can
we nd an operation O such that if P is a regular maniplex, then the symmetry
type graph of O(P) is precisely T ?
These questions give rise to voltage operations as a potential useful tool to nd 40
answers. Voltage operations dene a technique that uses voltage graphs to general-
ize the above mentioned classical operations, as well as many others, even in higher
dimensions (or ranks). These operations can also be dened for maniplexes. But
not only that, they can also be dened for premaniplexes, a more general concept
that includes both maniplexes and their symmetry type graphs. 45
It is important to remark that voltage graphs have been used before in con-
structing polytope-like structures. Notably, in [25] the authors use voltage graphs
to build 2-orbit n-maniplexes (for n ⩾ 4) with prescribed symmetry type graphs.
VOLTAGE OPERATIONS ON MANIPLEXES, POLYTOPES AND MAPS 3
In [4] the authors use voltage graphs, without explicitly mentioning them, to nd
generators and relations for the automorphism group of 3-orbit polytopes. In [21]
the second author of this manuscript exploited extensively the use of voltage graphs
to solve some relevant problems in the area, namely, problems 1 and 2 in [5].
5 We start this paper by giving some basic denitions that we shall need through-
out the manuscript. In Section 3 we give the denition of a voltage operation and
look in detail into the mix of two maniplexes (also called parallel product [30]) as a
voltage operation; we also study some connectivity properties of the voltage oper-
ations. In Section 4 we explore some examples of classical operations on polytopes
10 and polyhedra. In particular we show how pyramids, prisms, Wythoan opera-
tions, among others, can be seen as voltage operations. In Section 5 we describe
how automorphisms and voltage operations interact. We show that ifM is a mani-
plex with symmetry type T
O is a voltage operation, then every automorphism
and
of M induces an automorphism of O(M). That is, the automorphism group of M
15 is a subgroup of the automorphism group of O(M). Moreover, we show that the
symmetry type graph of O(M) with respect to the automorphism group of M
can be obtained by applying the voltage operation to T . In Section 6 we show
that voltage operations are closed under composition and we nd a simple way to
describe the composition of two voltage operations. Finally, in Section 7 we give
20 conditions for two voltage operations to be equivalent. We also characterize voltage
operations in terms of what they do to the ag graph of the universal polytope U
(introduced in [26] and dened in Section 2.2 as a maniplex). More precisely, in
Theorem 5.1 we prove that if O is an operation that assigns a premaniplex O(X ) to
each premaniplex X , then O is a voltage operation if and only if O(U/Γ) ∼
= O(U)/Γ
25 for every group Γ ⩽ Γ(U).
2. Preliminaries
2.1. Graphs. In this work we use the denition of graph used in [18], which
is slightly more general than the usual denition. A graph X is a quadruple
(D, V, I, (·)−1 ) where D and V are disjoint sets, I : D → V is a mapping and
30 (·)−1 : D → D is an involutory permutation of D. The set V is the set of vertices
of X , the set D is the set of darts of X . For a dart d, the vertex I(d) is the initial
vertex or starting point of d and d−1 is the inverse of d. The terminal vertex or
endpoint of d is the starting point of d−1 . The edges are the orbits of D under
−1
the action of (·) . If an edge consists only of a single dart, that is, a dart d that
35 satises d = d, then it is called a semiedge. A loop is an edge consisting of two
−1
darts whose initial vertex is the same. A link is an edge that is not a loop or a
semiedge. That is, an edge whose two darts are dierent and have dierent starting
points.
Usually a graph X is dened by its vertex set V (X) and the set E(X) of edges.
40 We can recover the set of darts as the set of formal pairs
for every dart d ∈ D(X). In other words, the initial vertex (resp. inverse) of the
image of a dart is the image of initial vertex (resp. inverse) of such dart. In this work
we shall evaluate graph homomorphisms (hence, isomorphisms and automorphism,
dened below, ) on the right, but any other function on the left. If v ∈ V (X) and 5
d ∈ D(X), we shall writevf and df instead of vfV and dfD . If both fV and fD are
bijective and f
−1
:= (fV−1 , fD
−1
) is also a graph homomorphism, then we say that f
(and f
−1
) is an isomorphism and that X and Y are isomorphic (and write X ∼ = Y ).
Naturally, for a graph X , an automorphism is an isomorphism f : X → X .
A path is a nite sequence W = (d1 , . . . , dk ) of darts such that the endpoint of 10
di is the starting point of di+1 for i ∈ {0, . . . , k − 1}. We usually omit commas and
parentheses and simply write W = d1 . . . dk . The number k is the length of W . The
starting point of d1 is the starting point of W ; we also say that W starts at I(d1 ).
Similarly, the endpoint of W is the endpoint of dk and we say that W ends at this
vertex. A single vertex is a path of length 0. A path is closed if its starting point 15
and its endpoint are the same vertex.
Notions such as valency of a vertex, cycle, subgraphs, connectivity, connected
components and trees extend naturally from the classic denition of graphs and
paths.
Given a path W = d1 . . . dk , an elementary (graph) move consists in inserting or 20
removing a pair of consecutive inverse darts at any point in the sequence d1 , . . . , dk .
If a path W′ can be obtained from a path W by applying a series of elementary
moves then we say that the paths are graph-homotopic (and write W ∼ W ′ ).
Clearly graph-homotopy is an equivalence relation and we often identify a path
with its homotopy class. Observe that if W is graph-homotopic to a path of length 25
0 then W must be closed.
W1 and W2 we say that they are compatible if the endpoint of W1
Given two paths
is the starting point of W2 . We can operate compatible paths by concatenation.
That is, if W1 = d1 . . . dk and W2 = a1 . . . aℓ then W1 W2 = d1 . . . dk a1 . . . aℓ . If
W1 ∼ W1′ and W2 ∼ W2′ then W1 W2 ∼ W1′ W2′ , which implies that we can operate 30
not only compatible paths but homotopy classes of compatible paths.
The fundamental groupoid of a graph X, denoted by Π(X) is the set of graph-
homotopy classes of X with the partial operation dened above. If u is a vertex in
X, then the fundamental group of X at u, denoted by Πu (X) is the set of graph-
homotopy classes of closed paths at u with concatenation as operation. Observe 35
that Πu (X) is actually a group and that if V is a path from v to u then Πv (X) =
V Πu (X)V −1 .
If X is a graph and G is a group, a voltage assignment is a function ξ : Π(X) → G
1
that satises ξ(W1 W2 ) = ξ(W2 )ξ(W1 ) for any two compatible paths W1 and V .
The group G is called the voltage group of ξ and the pair (X, ξ) is called a voltage 40
graph.
1
Usually a voltage assignment ξ is dened such that ξ(W1 W2 ) = ξ(W1 )ξ(W2 ) (cf. [18]). The
reason we do it the other way is because we are considering right actions, as is customary in the
polytopes and maniplexes literature.
VOLTAGE OPERATIONS ON MANIPLEXES, POLYTOPES AND MAPS 5
Since every path can be thought as the product of its darts, we can regard
a voltage assignment as a function ξ : D(X) → G such that for every dart d,
ξ(d−1 ) = ξ(d)−1 . The voltage of a path W = d1 , . . . , dk is simply ξ(dk ) · · · ξ(d1 ).
If (X, ξ) is a voltage graph, the derived graph is the graph X whose vertices and
ξ
5 darts are the elements in V (X) × G and D(X) × G, respectively. The initial vertex
−1
of a dart (d, g) is (I(d), g) and its inverse is (d , ξ(d)g). The adjacent vertices of
a given vertex (x, g) are the vertices (y, ξ(d)g) for each dart d starting at x and
ending at y . Equivalently, two vertices (x, g) and (y, h) are connected if there exists
−1
a dart d starting at x and ending at y whose voltage is hg .
10 If X is a graph and ξ : Π(X) → Γ and ζ : Π(X) → Γ are voltage assignments,
then ξ and ζ are equivalent if there is an isomorphism X → X such that the
ξ ζ
Xξ Xζ
(2.1)
X
where the arrows pointing to X are the projection to the rst coordinate, which is
a homomorphism.
15 The following Theorem is well known (see [18]):
M are in correspondence with the faces and sections of P. This close relation
between ag graphs of polytopes and maniplexes allows us to think of polytopes
in a graph-theoretical approach. In this paper, we slightly abuse language and
whenever we talk of a polytope P we actually refer to the ag graph of P. In
[9] Garza-Vargas and Hubard characterize when a maniplex is the ag graph of 5
a polytope. Finally, observe that the notions of ags, sections and faces extend
naturally to premaniplexes.
If X is a premaniplex, whenever we writex ∈ X we mean that that x is a ag
(vertex) in X.
x is a vertex in an n-premaniplex, and i ∈ {0, . . . , n − 1} we denote
If
by x the dart of color i whose starting point is x. We denote by x the i-adjacent 10
i i
The universal n-maniplex U n is the Cayley graph associated with the universal
Coxeter group C n . That is, the vertex set of U n is C n and for γ ∈ C n , the i-adjacent 40
n
vertex of γ is ρi γ . The maniplex U is in fact the ag graph of the universal
polytope {∞, . . . , ∞} introduced by Schulte in [26] (see [19, sect. 3D] [12, Theorem
5.2]). We omit the rank of both the universal Coxeter group and the universal
maniplex whenever it is implicit.
VOLTAGE OPERATIONS ON MANIPLEXES, POLYTOPES AND MAPS 7
of either adding or removing the same color two times at any two consecutive
w
positions (i.e. if v = ik , . . . , iℓ , j, j, iℓ−1 , . . . , i1 then x 7→ v x and
v
x 7→ w
x are
30 elementary moves) or swapping two non-consecutive colors in consecutive positions,
w
more precisely, if |iℓ −iℓ−1 | > 1, and u = ik , . . . , iℓ+1 , iℓ−1 , iℓ , iℓ−2 , . . . , i1 then x 7→
u
x is an elementary move. We say that two paths in a premaniplex are maniplex-
homotopic if we can turn one into the other by a nite sequence of elementary
(maniplex) moves. Observe that two paths in an premaniplex are homotopic if
35 their lifts to the universal maniplex U starting at a the same vertex also end at the
same vertex.
The homotopy class of a path in a premaniplex is uniquely determined by its
starting vertex and a monodromy of the universal maniplex. If w ∈ Mon(U n ) and
w
w = rik · · · ri1 w, then we denote by x the homotopy class of
is a word representing
40 the path ik ,ik−1 ,...,i1 x. Observe that any two words representing w yield homotopic
w
paths and if W1 , W2 ∈ x, then both paths end at wx.
Let X be a premaniplex and let Π(X ) be its fundamental groupoid. If ξ : Π(X ) →
Γ is a voltage assignment such that ξ(W ) is the identity in Γ whenever W is a path
8 VOLTAGE OPERATIONS ON MANIPLEXES, POLYTOPES AND MAPS
of length 4 alternating between two non-consecutive colors, we say that the pair
(X , ξ) is a voltage premaniplex. In other words, a voltage premaniplex is a voltage
graph where the voltage of the maniplex homotopy class of a path is well dened.
3. Voltage operations
i ̸∈ I
i∈I i∈I
given a vertexx, the dart of color i starting at x ends at d(ri )x. In other words,
Md , xi = d(ri )x. If M ∼
in = Md , we call each isomorphism φ : M → Md a
d-automorphism of M, and we say that M is d-automorphic.
d
The maniplex M can easily be seen as an (n, n)-voltage operation:
5 n
if (Y, η) = (1 , [d(r0 ), d(r1 ) . . . , d(rn−1 )]), then M ⋊η Y = M .
d
Classical examples of the above operation are the dual and the Petrial of an n-
maniplex X :
ri rj ri rj y = y,
Note now that since Y is a premaniplex, then ri rj ri rj y = y for every y ∈ Y , which
i,j,i,j
implies that (x, y) is closed if and only if η(i,j,i,j y) xes x. But η(i,j,i,j y) is
the identity, since (Y, η) is a voltage premaniplex. Therefore, we have the following
30 proposition.
we are thinking in terms of maniplex homotopy. In the same way, the notation
Π(Y) will denote the fundamental groupoid consisting of paths in Y considered up
to maniplex homotopy.
Although we have special interest on voltage operations that give rise to con-
nected premaniplexes, there are interesting examples of voltage operations in which 5
we (often) obtain disconnected objects.
A rooted premaniplex is a pair (X , x) where X is a connected premaniplex and
the root x is a vertex of X . If X is not connected, then (X , x) denotes the rooted
premaniplex with the connected component of x in X as its underlying premaniplex.
If(X , x) and (Y, y) are rooted premaniplexes and (Y, η) is a voltage operator, then 10
(X , x) ⋊η (Y, y) denotes the rooted premaniplex (X ⋊η Y, (x, y)).
Example 3.4. Let X be a n-premaniplex, −1 ⩽ k < ℓ ⩽ n and let (Y, η) =
(1ℓ−k−1 , [rk+1 , . . . , rℓ−1 ]). Then X ⋊η Y is a graph whose connected components
are the (k, ℓ)-sections of X . In particular, if k = −1 and X is a maniplex, then
X ⋊η Y determines the set of all ℓ-faces of X . Moreover, if y denotes the only vertex 15
in Y , for each x ∈ X , then (X , x) ⋊η (Y, y) is the ℓ-face of X that contains x.
µ the mixing voltage (for Y ) and we say that (Y, µ) is a mix operator. Then
call
X ⋊µ Y is the mix X ♢Y , and it is easy to see that this generalizes the same concept
previously dened for regular abstract polytopes. By rooting the premaniplexes,
this voltage operation generalizes the one for rooted maniplexes.
5 With our denition of the mix as a voltage operation we allow the resulting graph
to be disconnected. However, each connected component of X ⋊µ Y covers both
X and Y, so if at least one of them is simple (that is, a maniplex) the mentioned
components are maniplexes as well. Moreover, the connected component of X ♢Y
containing the vertex (x, y) is the smallest premaniplex that covers both (X , x) and
10 (Y, y) (cf. [5, Proposition 3.10]).
We can use this denition of the mix to nd the smallest cover of a maniplex
satisfying some property, as we shall see in the next example.
Example 3.5. Let I ⊆ {0, 1, . . . , n − 1}. We denote by 2nI the n-premaniplex with
2 vertices with semiedges of colors in I at each vertex and links of colors not in I
15 between the 2 vertices (see Figure 2b). We can use these premaniplexes to describe
certain properties of maniplexes: orientable n-maniplexes (those that are bipartite)
are those that cover 2n∅ and vertex-bipartite maniplexes (those whose 1-skeleton is
bipartite) are those that cover 2n{1,2,...,n−1} , for example. In general, if X covers 2nI
it means that there is a coloring of the vertices (or ags) of X with two colors such
20 that i-adjacent ags are of the same color if and only if i ∈ I . See [17, 25] for a
detailed discussion on ag colorings.
If a maniplex X does not cover 2nI , then X ♢2nI is a maniplex that covers X
n
but also covers 2I . We call this the X double cover of with respect to
I . If X is
n
non-orientable, X ♢2∅ is the so called orientable double cover of
X.
25 n n n
Note that if X does cover 2I , then X ♢2I = X ⋊µ 2I (with µ the mixing voltage)
consists of two isomorphic copies of X , but the ags of these copies will be colored
n n
by the vertices of 2I . More precisely, if we colored the vertices of 2I with white
n
and black, then each vertex (ag) of X ♢2I is colored white or black according to
n
its second coordinate. So X ♢2I consists of two copies of X but together with an
30 I -compatible coloring, and the two copies have opposite colorings. In particular,
if X is orientable and I = ∅, the two copies of X ♢2n∅ are mirror images; if X is a
chiral maniplex, then X ⋊µ 2n∅ consists of its two enantiomorphic forms.
Before ending this section let us note that the mixing voltages are the only
voltages such that the product is naturally commutative. More precisely:
35 Proposition 3.6. Let (X , ηX ) and (Y, ηY ) be (n, n)-voltage operators. Then the
function (x, y) 7→ (y, x) is an isomorphism between X ⋊ηY Y and Y ⋊ηX X if and
only if both ηX and ηY are mixing voltages.
Proof. The alleged function is an isomorphism if and only if it maps (x, y)i =
i i i i i
(ηY ( y)x, y ) to (y, x) = (ηX ( x)y, x ), for all x ∈ X , y ∈ Y and i ∈ {0, 1, . . . , n−1}.
40 i
This means that ηY ( y) = ri and ri = ηX ( x).
i
□
Note that X ⋊η Y Y and Y ⋊ηX X may be isomorphic even if ηX and ηY are not
mixing voltages, for example, if (X , ηX ) is the orientable double cover and (Y, ηY )
is the snub (or, of course, if (X , ηX ) = (Y, ηY )).
X is connected X ⋊η Y is connected as well (in the context of [6], these are called
fully stratied operations ). We rst analyze when we can nd a path between two
vertices of X ⋊η Y and use this to determine when a voltage operator preserves
connectivity.
Lemma 3.7. Let (x, y) and (x′ , y ′ ) be two vertices of X ⋊η Y . There is a path in 5
X ⋊η Y that starts at (x, y) and ends at (x′ , y ′ ) if and only if there exists a path W
in Y that connects y with y ′ and such that η(W )x = x′ .
Proof. Assume there is a path ik ,...,i1 (x, y) that ends at (x′ , y ′ ). By Remark 3.2 the
ik ,...,i1
path (x, y) ends at (η(W )x, rik · · · ri1 y), where W denotes the path ik ,...,i1 y .
Hence, η(W )x = x and rik · · · ri1 y = y . Moreover, rik · · · ri1 y is precisely the 10
′ ′
′
endpoint of W , which implies that W connects y with y .
′
Conversely, assume that there exists a path W that connects y with y and
′ ik ,...,i1
satises that η(W )x = x . Since W starts at y , it can be written as W = y , for
′
some ik , . . . , i1 ∈ {0, . . . , m − 1}. This implies that y = rik · · · ri1 y . By Remark 3.2,
the path k
i ,...,i1
(x, y) ends at (η(W )x, rik · · · ri1 y) = (x′ , y ′ ). 15
□
Proposition 3.8. The graph X ⋊η Y is connected if and only if Y is connected and
η(Πy0 (Y)) acts transitively on X for some vertex y0 ∈ Y .
Proof. We start by assuming that X ⋊η Y is connected, and let x, x′ ∈ X . First note
that X ⋊η Y covers Y , so Y must be connected as well. Now we want to prove that 20
η(Πy0 (Y)) acts transitively on X . Since X ⋊η Y is connected, there is a path from
(x, y0 ) to (x′ , y0 ). By Lemma 3.7 there exists W ∈ Πy0 (Y) such that η(W )x = x′ .
′ y0
Since x and x were arbitrary, we have proved that η(Π (Y)) acts transitively on
X.
Now we assume that Y is connected and η(Π (Y)) acts transitively on X . We 25
y0
Here, we shall see some examples of operations on maps and polytopes that
can be seen as voltage operations. In fact, if we have an operation on polytopes
(or maps) that can be seen as a voltage operation, by dening the corresponding
5 voltage operator we have dened the operation on maniplexes and premaniplexes.
We show here that the Wythoan constructions, including the snub, all prisms
and pyramids over polytopes, the trapezotope and the k -bubbles can be seen as
voltage operations (each of them will be properly dened). Moreover, we also
study the operations M 7→ 2M and M 7→ 2̂M , and see that although there is no
10 voltage operator that will give us these operations, whenever M is regular, there
is a way to see the construction 2̂M can be described as a voltage operation that
depends on each M.
of ot(P) are the ags of P and an edge of the form Φ \ {Φi } joins the vertices Φ
14 VOLTAGE OPERATIONS ON MANIPLEXES, POLYTOPES AND MAPS
r1 r0
r1 r1
r0 r2 r2 r2
r1
r1 r2 r0 r2
r0 r0 r1 r1
r1 r0 r2
r0 r1
r0 r2
r2
(e) A = {ρ0 , ρ1 , ρ2 }
and Φi . This means that the 1-skeleton of ot(P) is the ag graph of P . Another
way to construct ot(P) is as the colorful polytope of the ag graph of P [1].
The omnitruncation in rank n can be thought as a voltage operation. To see this,
use the following construction for the voltage operator. Let S be the ag graph of
an (n − 1)-simplex. First, change the color of each edge by increasing it by 1 (i.e. if 5
two ags of S are i-adjacent, they will be joined by an edge of color i + 1). All these
edges will have the identity voltage. Note that now each vertex has an edge of color
i, for i ∈ {1, 2, . . . , n − 1}. Now, label each connected component of the graph with
VOLTAGE OPERATIONS ON MANIPLEXES, POLYTOPES AND MAPS 15
r0 r2
r0 r1 r2 r 1
r0 r1 r2 r1
denoted by Pri(P) is dened as the poset ((P \ P−1 ) × Λ) ∪ {F−1 } where Λ is the
poset {{λ0 }, {λ1 }, {λ0 , λ1 }} ordered by inclusion and F−1 is the unique minimum
25 element of the prism.
Both the prism and the pyramid over an n-polytope are (n + 1)-polytopes. They
are particular cases of products of polytopes, in the sense of [10]: the pyramid is
the join product by a vertex, while the prism is the direct product by an edge.
One can see that each ag Ψ in Pyr(P) is of the form
Ψ = {(Φ−1 , 0), (Φ0 , 0), . . . , (Φt , 0), (Φt , 1), (Φt+1 , 1), . . . , (Φn , 1)},
30 where Φ is a ag in P and t ∈ {−1, 0, 1, . . . , n}. Hence, we identify the set of ags
of Pyr(P) with F(P) × {−1, 0, . . . , n}.
16 VOLTAGE OPERATIONS ON MANIPLEXES, POLYTOPES AND MAPS
In a similar way, one can identify the set of ags of Pri(P) with F ×{0, 1, . . . , n}×
{λ0 , λ1 }, where the last coordinate tells us if the 0-face is of the form (Φ0 , {λ0 }) or
(Φ0 , {λ1 }).
Let us rst observe the pyramid. The i-adjacencies of the ags (see [10]), for
i ∈ {0, 1, . . . , n}, are given by 5
i
(Φ , t) if 0 ⩽ i < t,
(Φ, t − 1) if i = t,
(Φ, t)i =
(Φ, t + 1) if i = t + 1,
i−1
(Φ , t) if t + 1 < i.
Let (Y, η) be the (n, n + 1)-voltage operator whose vertices are the numbers
{−1, 0, . . . , n} with the i-adjacencies (for i ∈ {0, 1, . . . , n}) given by:
t
if i < t or t + 1 < i,
i
t = t − 1 if i = t,
t + 1 if i = t + 1,
ri
if i < t,
i
ξ( t) = 1 if i = t or i = t + 1,
ri−1 if t + 1 < i.
(Φi , t, λ) if i < t,
′
(Φ, t, λ ) 0 = i = t,
if
i
(Φ, t, λ) = (Φ, t − 1, λ) if 0 < i = t,
(Φ, t + 1, λ) if i = t + 1,
(Φi−1 , t, λ)
if t + 1 < i,
1 2 t t+1 n−1 n
1 2 t t+1 n−1 n
r1
r1 r0
(0, 0, 1)
r0
(0, 1, 0) r1
r1 r0
(0, 0, 0) (1, 0, 0)
r0
the setF(P)×Zn+12 . When we denote a ag of Trp(P) by (Φ, v), the i-th coordinate
of the vector v tells us if the faces (Φ, v)i = (Fi , Gi ) and (Φ, v)i−1 = (Fi−1 , Gi−1 )
dier in the rst coordinate (when vi = 0) or in the second one (when vi = 1).
Using this natural correspondence, one can prove that the i-adjacent ag of the
ag (Φ, v) is: 5
(Φ, v + e0 )
if i = 0,
where j is dened the same way as before. One can conrm that the voltage 10
operator(Y, η) satises that for every polytope P, the ag graph of Trp(P) is
G(P) ⋊η Y .
The k-bubble. The k -bubble operation was introduced by Helfand in [13] as a gen-
eralization of the truncation of the vertices.
VOLTAGE OPERATIONS ON MANIPLEXES, POLYTOPES AND MAPS 19
ri ri ri ri ri ri
rk+2
ri ri+1 ri
ri ri ri+1 ri+1
The order in [P]k is dened by the following rules: Let H and H ′ be faces in P
with ranks dierent than k, let F be a k -face of P and let G and G′ be faces of P
5 properly containing F.
• H < H in [P]k if and only if H < H ′ in P .
′
(ri Φ, ℓ) if i ⩽ k
(r Φ, ℓ) if k + 1 ⩽ i ⩽ ℓ − 2,
i+1
i
(Φ, ℓ) = (Φ, ℓ − 1) if k + 1 ⩽ i = ℓ − 1,
(Φ, ℓ + 1) if i = ℓ,
(r Φ, ℓ)
i if i ⩾ ℓ + 1.
With this information we can see that the k -bubble is described by the voltage
15 operator in Figure 8.
(
i (Ψi , v) if i < n,
(Ψ, v) =
(Ψ, v + ek ) i = n, and the facet of Ψ is labeled with k.
where ek ∈ Zm
2 is the vector that has 0 in all its entries except the k -th one.
These operations on maniplexes are very interesting. In particular in the context
of voltage operations, Example 5.3 will show that one cannot nd a premaniplex
Y that satises that 2M = M ⋊η Y , for every maniplex M (not even if we x the 10
rank of M, or ask M to be regular). However we shall see that if M is a regular
premaniplex, then there exists a premaniplex YM and a voltage assignment η such
M
that 2̂ = M ⋊η YM .
Let M be a regular n-premaniplex and let ρ0 , ρ1 , . . . , ρn−1 be the distinguished
generators of Γ(M) with respect to the base ag Φ. We dene the premaniplex 15
Y := YM as follows. The set of vertices of Y is the set Zm 2 . Given v ∈ Y the
entries of v correspond to facets of M. Assume that the facets of M are labeled
with the set {1, . . . , m}, as above, in such a way that the facet corresponding to
the base ag is labeled with 1. Notice that every element in Γ(M) permutes the
m
facets of M, which in turn induces an action of Γ(M) on Z2 by permutation of 20
m
coordinates. More precisely, for α ∈ Γ(M) and v = (v1 , . . . vm ) ∈ Z2 , let vα denote
the vector (v1α−1 , . . . , vmα−1 ), that is, the vector resulting from v after permuting
the coordinates according to α.
The adjacencies in Y are given as follows: for i ∈ {0, 1, . . . , n − 1} there is an
i
edge of color i between v and v := vρi . Further, there is an edge of color n between 25
v and v + e1 . We now dene the voltage assignment on Y as:
(
i ri ifi < n,
η( v) =
1 i = n.
Question 4.1. Is it true that for every maniplex M, there exists a voltage operator
(YM , η) such that 2̂M ∼
= M ⋊η YM ?
10 For example, we have not been able to determine if there exists such voltage
operator for the case when M is the pyramid over a digon. If the answer for the
above question is negative, then we could ask:
5. Automorphisms
In this section we will study the interplay between the automorphism groups of
a premaniplex and the resulting premaniplex after a voltage operation, as well as 5
their symmetry type graphs.
Let X (Y, η) be a voltage operator. Observe that Γ(X )
be a premaniplex and let
acts faithfully as a group of automorphisms on X ⋊η Y . Indeed, if γ ∈ Γ(X ), then
the mapping γ : (x, y) 7→ (xγ, y) induces an automorphism of X ⋊η Y . To see this,
just note that γ commutes with the monodromies 10
ri ((x, y)γ) = ri (xγ, y)
= η(i y)xγ, ri y
= η(i y)x, ri y γ
(φ(x), y).
Now, using the fact that φ is a homomorphism we simply check that
= φ(η(i y)x), ri y
= φ′ η(i y)x, ri y
Example 5.3. In this example we will prove that there is no (n, n + 1)-voltage
operator (Y, η) such that 2M = M ⋊η Y for every n-maniplex M (as was hinted
35 in Question 4.1). For a 2-gon {2}, the polytope 2{2} is the quadrangular dihedron
{4, 2} (the map on the sphere consisting of two squared faces sharing all the vertices
{4}
and edges). For the square {4}, we have that 2 is the toroidal map {4, 4}(4,0)
(a 4 × 4 grid on the torus). We can get a 2-gon from a square by taking its
quotient by the cyclic group of order two generated by the half-turn around the
40 center, that is {2} = {4}/Z2 . However, it is not possible to get the quadrangular
dihedron by taking a quotient of {4, 4}(4,0) by any group of order 2. In other words
2{4} /Z2 ≇ 2{4}/Z2 . Hence, Theorem 5.1 is not satised for the operation 2M . Thus,
there is no voltage operator (Y, η) such that 2M = M ⋊η Y for every maniplex M.
The previous example uses Theorem 5.1 to prove that there does not exist a
45 voltage operation (Y, η) such that 2M ∼= M ⋊η Y for all maniplexes M, in other
words, that M 7→ 2M is not a voltage operation. Observe that this fact can also
24 VOLTAGE OPERATIONS ON MANIPLEXES, POLYTOPES AND MAPS
Theorem 5.4. Let (X , ξ) be a voltage premaniplex with voltage group Γ and let 20
(Y, η) be a voltage operator.
Dene θ = θ(η, ξ) : Π(X ⋊η Y) → Γ as follows:
ω
(5.1) θ(ω (x, y)) := ξ(η( y)
x),
for every ω ∈ Mon(U).
Then (X ⋊η Y)θ is isomorphic to X ξ ⋊η Y .
Proof. Start by noticing that, by denition, ri (x, γ) = (ri x, ξ(i x)γ) and similarly
i
ri ((x, y), γ) = (ri (x, y), θ( (x, y))γ). Now, dene the function
φ : (X ⋊η Y)θ → X ξ ⋊η Y
given by
φ((x, y), γ) := ((x, γ), y).
We shall show that φ is a premaniplex isomorphism. 25
Since both Xξ and (X ⋊η Y)θ have Γ as voltage group, φ is a bijection. We
should see next that φ preserves i-adjacencies for i ∈ {0, 1, . . . , m − 1}. In fact:
= ri ((x, γ), y)
= ri φ ((x, y), γ) .
VOLTAGE OPERATIONS ON MANIPLEXES, POLYTOPES AND MAPS 25
□
Example 5.5. It is not dicult to see that the antiprism of a q -gon can be obtained
by taking the medial of the pyramid over a q -gon. Thus, we shall use Theorem 5.4 to
recover the antiprism of a q -gon as a voltage maniplex. To do so, we rst construct
5 the pyramid over a q -gon as the derived graph of a voltage premaniplex and then
apply the medial operator (as a voltage operator).
Note that the automorphism group of a q -gonal pyramid coincides with the
automorphism group of its base, which is the dihedral group
30 Corollary 5.6. Let X be a regular n-premaniplex with automorphism group ⟨ρ0 , . . . , ρn−1 ⟩
and let (Y, η) be a (n, m)-voltage operator. Then X ⋊η Y is isomorphic to the de-
rived graph Y ν , where ν : Π(Y) → Γ(X ) is the voltage assignment obtained from η
by replacing each ri with ρi .
Proof. Since X is regular, it is isomorphic to the derived graph where, if x (1n )ξ
35 is the only vertex of
n i
1 , ξ( x) = ρi . Then Theorem 5.4 tells us that X ⋊η Y is
n θ ω η(ω y
isomorphic to (1 ⋊η Y) with θ( (x, y)) = ξ( )x). This means precisely that θ
n
replaces each occurrence of ri in η by the voltage of the semiedge of color i in 1 ,
n
but this is exactly ρi . By applying the natural isomorphism 1 ⋊η Y → Y we get
the desired result. □
r0 r1 ρ0 ρ1 ρ0 ρ0
(Y, η) (X ⋊η Y, θ)
r1 r2 ρ1 ρ1 ρ0 ρ1
ρ0 ρ0
ρ0
a b c d
ρ1 ρ1
ρ1
(X , ξ)
a maniplex, so one must be careful with this fact when composing operations, as
the result of a voltage operation can be disconnected.
It is interesting to note that in fact the composition of two voltage operations
can be written as a new voltage operation. In this section we describe how to do
this by using the operator θ dened in Theorem 5.4, but instead of using an arbi- 5
trary voltage premaniplex (X , ξ) and a voltage operator (Y, η) we use two voltage
operators (Y1 , η1 ) and (Y2 , η2 ).
Theorem 6.1. Let X be an n-premaniplex, (Y1 , η1 ) an (n, m)-voltage operator and
(Y2 , η2 ) a (m, ℓ)-voltage operator. Then
(X ⋊η1 Y1 ) ⋊η2 Y2 ∼
= X ⋊θ (Y1 ⋊η2 Y2 ),
where θ = θ(η2 , η1 ) is dened as in Equation (5.1) .
Proof. We dene φ : (X ⋊η1 Y1 ) ⋊η2 Y2 → X ⋊θ (Y1 ⋊η2 Y2 ) in the natural way,
that is,
= ri (x, (y1 , y2 ))
= ri φ ((x, y1 ), y2 )
□
VOLTAGE OPERATIONS ON MANIPLEXES, POLYTOPES AND MAPS 27
The above theorem can be applied in dierent contexts, we give some examples
here. Let (1m , d) be the duality operator and consider a (n, m)-voltage operator
(Y, η). If X is a premaniplex, then (X ⋊η Y) ⋊d 1m (X ⋊η Y).
is the dual of
m
Theorem 6.1 tells us that this dual is in fact isomorphic to X ⋊θ (Y ⋊d 1 ). Thus,
5 (Y ⋊d 1m , θ) is the dual of Y, where the voltages of the darts are the same as the
ones of the darts of the dual color in (Y, η). In other words, if (Y ∗ , η ∗ ) denotes
the voltage operator we get by recoloring the darts of color i in Y with the color
n − 1 − i, then the dual of X ⋊η Y is X ⋊η ∗ Y ∗ .
More generally, Theorem 6.1 lets us dene the composition of two operators.
10 If (Y1 , η1 ) is an (n, m)-operator and (Y2 , η2 ) is an (m, ℓ)-operator we dene the
composition of (Y1 , η1 ) with (Y2 , η2 ) as (Y1 ⋊η2 Y2 , θ(η1 , η2 )) and denote it by
(Y1 , η1 ) ◦ (Y2 , η2 ). Theorem 6.1 tells us that (Y1 , η1 ) ◦ (Y2 , η2 ) is an (n, ℓ)-operator
and that the composition of operators is associative. This allows us to dene a new
category: recall that pMpxn denotes the class of all premaniplexes of rankn and
15 let pMpx be the category whose objects are the classes pMpxn
n ⩾ 1, and
with
n
whose arrows are voltage operators. An (n, m)-operator is an arrow from pMpx to
pMpxm and the composition of arrows is dened as above. The neutral element at
the object pMpxn is the arrow (1n , µ) where µ is the mixing voltage, and the iso-
morphisms are precisely the voltage operators described in Example 3.1. It might
20 be also interesting to study the analogous category obtained by considering rooted
voltage operations.
Observe that the snub operation seems to act dierently on orientable maniplexes
than on non-orientable ones. However, the result of applying the snub operation
to a non-orientable maniplex M is one of the connected components of doing the
25 same operation to the orientable double cover of M. This phenomenon is easy to
understand with the following results:
Proof. Because of Theorem 2.1, we may assume without loss of generality that
η(W ) ∈ η(Πy2 (Y2 )) for every path W ∈ Π(Y2 ), and thus η(W ) xes y1 .
First we notice that the induced (colored) graph of Y1 ⋊η Y2 with vertex set
35 {(y1 , y) : y ∈ Y2 } forms an isomorphic copy of Y2 . This is easy to see since
the i-adjacent ag to Y1 ⋊η Y2 ) is
(y1 , y) (in (η(i y)y1 , y i ) = (y1 , y i ), because, by
assumption, the voltages of paths in Y2 x y1 .
Next we use Theorem 6.1 to see (X ⋊µ Y1 ) ⋊η Y2 as X ⋊θ (Y1 ⋊η Y2 ) with
θ = θ(η, µ). This means that we can consider the i-adjacent ag to ((x, y1 ), y) in
40 (X ⋊µ Y1 ) ⋊η Y2 as the i-adjacent ag to (x, (y1 , y)) in X ⋊θ (Y1 ⋊η Y2 ); we write
((x, y1 ), y) ↔ (x, (y1 , y)) to denote that these two points are in correspondence
under the isomorphism. On one hand, observe that by denition of θ , for any
ω
y ∈ Y2 and any monodromy ω , we have that θ(ω (y1 , y)) is µ(η( y) y1 ). On the other
ω
η( y)
hand, by denition of µ we have that µ( y1 ) = η(ω y).
28 VOLTAGE OPERATIONS ON MANIPLEXES, POLYTOPES AND MAPS
Hence,
i i
(x, y1 ), y ↔ x, (y1 , y)
= θ(i (y1 , y))x, (yi , y)i
i
= µ(η( y)
y1 )x, (y1 , y)i
The proof of Theorem 7.1 is based on the fact that since U is regular, its
monodromy group acts regularly on its vertices (ags). We can generalize The-
orem 7.1 as follows. Let X be a regular premaniplex and let (Y, η) be a voltage
operator. SX = {ω ∈ Mon(U) : ωx = x for all x ∈ X }, that is, the ker-
If
5 nel of the projection Mon(U) → Mon(X ), then Mon(X ) ∼ = Mon(U)/SX . Let
ηX : Π(Y) → Mon(X ) be the voltage assignment on Y we get by reducing η to
Mon(X ), that is ηX (W ) := η(W )SX . Then by using the exact same argument as
η
in Theorem 7.1 we can see that X ⋊η Y is isomorphic to Y X .
An immediate consequence of Theorems 5.1 and 7.1 is that operators (Y, η1 ) and
10 (Y, η2 ) with η1 and η2 equivalent voltages (that is, with isomorphic derived graphs
as in Equation (2.1)) yield equivalent voltage operations. More precisely,
Proposition 7.2. Let (Y, η1 ) and (Y, η2 ) be two (n, m)-operators with equivalent
voltages η1 and η2 . Then for every n-premaniplex X ,
X ⋊η 1 Y ∼
= X ⋊η2 Y.
Moreover, there is an isomorphism X ⋊η1 Y → X ⋊η2 Y such that the following
15 diagram commutes:
X ⋊η 1 Y X ⋊η 2 Y
(7.1)
X ⋊η1 Y ∼
= (U/Γ) ⋊η1 Y
∼
= (U ⋊η Y) /Γ 1
∼
= Y η1 /Γ
∼
= Y η2 /Γ
∼
= (U ⋊η2 Y) /Γ
∼
= (U/Γ) ⋊η Y 2
∼
= X ⋊η2 Y.
If we start with a vertex (x, y) in X ⋊η 1 Y and apply the natural isomorphisms
between consecutive terms in the equation above, we get the following sequence,
where Φ is a base ag of U:
(x, y) 7→ (ΨΓ, y) for some Ψ∈U
7→ (Ψ, y)Γ
7→ (y, ω)Γ where ω ∈ Mon(U) is such that Ψ = ωΦ
7→ (y, ω̃)Γ for some ω̃ ∈ Mon U (since η1 is equivalent to η2 )
7→ (Ψ̃, y)Γ where Ψ̃ = ω̃Φ
7→ (Ψ̃Γ, y)
7→ (x̃, y) for some x̃ ∈ X .
20 We can see that this isomorphism makes the diagram in Equation (7.1) commute.
□
30 VOLTAGE OPERATIONS ON MANIPLEXES, POLYTOPES AND MAPS
= O (U/ Γ(U)) ∼
O(U)/ Γ(U) ∼ = O(1n ) = Y,
which implies that there exists a voltage assignment ∼ Mon(U)
η : Π(Y) → Γ(U) =
such that Yη ∼ = O(U). The pair (Y, η) denes a voltage operator and by Theo-
rem 7.1, U ⋊η Y ∼= Yη ∼ = O(U).
Finally, if X is a premaniplex and Γ ⩽ Γ(U) is such that X ∼
= U/Γ, then 15
O(X ) =∼ O(U/Γ) ∼ = O(U)/Γ ∼= (U ⋊η Y)/Γ ∼= (U/Γ) ⋊η Y ∼= X ⋊η Y. □
We often come across operations that are well dened for some family F of pre-
maniplexes (for example, for convex polytopes) but such that it is not immediately
evident how to generalize them for all premaniplexes. One may ask if it is possible
to extend operations of this kind to all premaniplexes (of the given rank) in such
a way that the operation is a voltage operation. As an example, in Section 4, we 20
have found voltage operations that extend the Wythoan operations, the k -bubble
and the trapezotope operation to all premaniplexes. Theorem 7.3 and Corollary 5.6
answer this question and nd the corresponding voltage operation, when possible.
The idea is as follows: suppose there is some regular premaniplexP and an opera-
tion O O(P). According to the proof of Theorem 7.3, if 25
such that we already know
O is indeed a voltage operation X 7→ X ⋊η Y , then Y should be O(1n ), but due to
Theorem 5.1 Y must coincide with O(P)/ Γ(P). Now let ν : Π(Y) → Γ(P) be the
ν
voltage assignment such that O(P) is isomorphic to Y . Corollary 5.6 tells us that
ν is obtained by replacing each occurrence of ri by ρi in the voltage assignment
η : Π(Y) → Mon(U). So we can recover η by replacing every instance of ρi by ri 30
in ν . Note that, in general, such replacement is not well dened unless P is the
universal polytope U , however with some intuition we can nd the right way to
do it for the natural occurring operations. However, observe that one can do this
without knowing if O was in fact a voltage operation. To know if O is a voltage
operation or not, one must see that O(X ) = X ⋊η Y for X in the family F . In fact, 35
if for some X ∈F we observe that O(X ) ̸= X ⋊η Y , then O cannot be seen as a
voltage operation.
We have seen that voltage operations generalize classical operations on maps and
polytopes and allow us to dene such classical operations on premaniplexes. 40
One can see that voltage operations naturally generalize to hypertopes (thin
residually connected geometries), complexes, and their quotients. Following [31],
VOLTAGE OPERATIONS ON MANIPLEXES, POLYTOPES AND MAPS 31
Problem 1. Given a premaniplex T, does there exist a polytope (or maniplex) such
that its symmetry type graph (with respect to the full automorphism group) is T?
A particular example of this problem was the question posted in the early 1990's
15 by Schulte and Weiss of whether or not there exist chiral polytopes of all ranks
(which was solved by Pellicer in 2010 [24]). Of course, one can generalize Problem 1
to hypertopes.
Problem 2. Given a precomplex T, does there exist a hypertope (or complex) such
that its symmetry type graph (with respect to the full automorphism group) is T?
20 One thing that one tries to do when dealing the above question is to use voltage
assignments on T to lift to a maniplex M. In [21] the voltage assignments that give
a polytope as the derived graph are characterized. Of course the voltage group acts
by automorphisms on M and the quotient of M by it is precisely T. However, M
can have (and often does) symmetries not coming from the voltage group. In the
25 context of voltage operations, we have shown that given a voltage operator(Y, η)
and a premaniplex X , all automorphisms of X act as automorphisms of X ⋊η Y .
However, again, X ⋊η Y might have extra symmetry (for example, in the case when
one applies the medial operation to a self-dual map). It is natural to ask when is
it true that automorphisms of Y also lift to X ⋊η Y , and when all automorphisms
30 of X ⋊η Y come either from X or from (Y, η). We refer to [15] for more details
about these questions. Answering, at least partially, these questions might be of
great help for solving Problem 1 (at least partially).
Problem 3. Give necessary conditions on (Y, η) so that one can compute the sym-
metry type graph of X ⋊η Y with respect to its full automorphism group in terms
35 of X and (Y, η).
If one's interest is in polytopes rather that in (pre)maniplexes, one can ask when
a voltage operation preserves polytopality, though we might not care if the result
of the operation is not connected. More precisely,
Problem 4. Give necessary and sucient conditions on (Y, η) for (P, Φ) ⋊η (Y, y)
40 to be a polytope, for all rooted polytopes (P, Φ) and y ∈ Y.
Similarly,
Problem 5. Give necessary and sucient conditions on (Y, η) for (H, Φ) ⋊η (Y, y)
to be a hypertope, for all rooted hypertopes (H, Φ) and y ∈ Y.
32 REFERENCES
Acknowledgements
The authors thank the nancial support of CONACyT grant A1-S-21678. This
paper was completed while the second author held a Post Doctoral Research Asso-
ciate position in the Department of Mathematics at Northeastern University and the
third author was supported by the Post Doctoral Scholarship Program at UNAM 5
(DGAPA), Mexico.
References
[1] Gabriela Araujo-Pardo, Isabel Hubard, Deborah Oliveros, and Egon Schulte.
Colorful polytopes and graphs. en. In: Israel Journal of Mathematics 195.2
(June 2013), pp. 647675. issn: 1565-8511. doi:10 . 1007 / s11856 - 012 - 10
0136-7. url: https://doi.org/10.1007/s11856-012-0136-7 (visited on
10/19/2021).
[2] H. S. M. Coxeter. Regular polytopes. Third. Dover Publications, Inc., New
York, 1973, pp. xiv+321. doi: 10.2307/1573335.
[3] Harold Scott Macdonald Coxeter and Harold Scott Macdonald Coxeter. The 15
beauty of geometry: Twelve essays. Courier Corporation, 1999.
[4] Gabe Cunningham, María Del Río-Francos, Isabel Hubard, and Micael Toledo.
Symmetry type graphs of polytopes and maniplexes. In: Ann. Comb. 19.2
(2015), pp. 243268. issn: 0218-0006. doi:10.1007/s00026- 015- 0263- z.
url: https://doi.org/10.1007/s00026-015-0263-z. 20
[5] Gabe Cunningham and Daniel Pellicer. Open problems on k -orbit polytopes.
In: Discrete Math. 341.6 (2018), pp. 16451661. issn: 0012-365X. doi: 10.
1016/j.disc.2018.03.004. url: https://doi.org/10.1016/j.disc.
2018.03.004.
[6] Gabe Cunningham, Daniel Pellicer, and Gordon Williams. Stratied opera- 25
tions on maniplexes. To appear.
[7] L. Danzer. Regular incidence-complexes and dimensionally unbounded se-
quences of such, I. In: North-Holland Mathematics Studies. Vol. 87. Elsevier,
1984, pp. 115127.
[8] Ian Douglas, Isabel Hubard, Daniel Pellicer, and Steve Wilson. The twist 30
operator on maniplexes. In: Geometry and Symmetry Conference. Springer.
2015, pp. 127145.
[9] Jorge Garza-Vargas and Isabel Hubard. Polytopality of maniplexes. In: Dis-
crete Math. 10.1016/j.
341.7 (2018), pp. 20682079. issn: 0012-365X. doi:
disc.2018.02.017. url: https://doi.org/10.1016/j.disc.2018.02. 35
017.
[10] Ian Gleason and Isabel Hubard. Products of abstract polytopes. In: Journal
of Combinatorial Theory, Series A 157 (July 2018), pp. 287320. doi: 10.
1016/j.jcta.2018.02.002.
[11] Ian Gleason and Isabel Hubard. The antiprism of an abstract polytope. 40
en. In: ARS MATHEMATICA CONTEMPORANEA 0 (Aug. 2021). issn:
10 . 26493 / 1855 - 3974 . 2584 . 68d. url: https : / / amc -
1855-3974. doi:
journal.eu/index.php/amc/article/view/2584 (visited on 10/28/2021).
[12] M. I. Hartley. All polytopes are quotients, and isomorphic polytopes are
quotients by conjugate subgroups. In: Discrete Comput. Geom. 21.2 (1999), 45
pp. 289298. issn: 0179-5376. doi: 10.1007/PL00009422. url: https://
doi.org/10.1007/PL00009422.
REFERENCES 33
[28] Egon Schulte and Abigail Williams. Wythoan Skeletal Polyhedra in Ordi-
nary Space, I. In: Discrete & Computational Geometry 56.3 (2016), pp. 657
692.
[29] Andrew Vince. Combinatorial maps. In: Journal of Combinatorial The-
ory. Series B 34.1 (1983), pp. 121. issn: 0095-8956. doi: 10.1016/0095- 5
8956(83)90002-3.
[30] Stephen E. Wilson. Parallel products in groups and maps. In: J. Algebra
10.1006/jabr.1994.1200.
167.3 (1994), pp. 539546. issn: 0021-8693. doi:
url: https://doi.org/10.1006/jabr.1994.1200.
[31] Steve Wilson. Maniplexes: Part 1: maps, polytopes, symmetry and opera- 10
tors. In: Symmetry 4.2 (2012), pp. 265275. issn: 2073-8994. doi: 10.3390/
sym4020265. url: https://doi.org/10.3390/sym4020265.