Conference PaperPDF Available

Block Representation of Reversible Causal Graph Dynamics

Authors:

Abstract and Figures

Causal Graph Dynamics extend Cellular Automata to arbitrary, bounded-degree, time-varying graphs. The whole graph evolves in discrete time steps, and this global evolution is required to have a number of physics-like symmetries: shift-invariance (it acts everywhere the same) and causality (information has a bounded speed of propagation). We study a further physics-like symmetry, namely reversibility. More precisely, we show that Reversible Causal Graph Dynamics can be represented as finite-depth circuits of local reversible gates.
Content may be subject to copyright.
Reversible causal graph dynamics
Pablo Arrighi1, Simon Martiel2, and Simon Perdrix3
1Aix-Marseille Univ., LIF, F-13288 Marseille Cedex 9, France
pablo.arrighi@univ-amu.fr,
2Univ. Nice-Sophia Antipolis, I3S, 06900 Sophia Antipolis, France
martiel@i3s.unice.fr
3CNRS, LORIA, Inria Project Team CARTE, Univ. de Lorraine, Nancy, France
simon.perdrix@loria.fr
Abstract. Causal Graph Dynamics extend Cellular Automata to arbi-
trary, bounded-degree, time-varying graphs. The whole graph evolves in
discrete time steps, and this global evolution is required to have a num-
ber of physics-like symmetries: shift-invariance (it acts everywhere the
same) and causality (information has a bounded speed of propagation).
We add a further physics-like symmetry, namely reversibility.
Keywords. Bijective, invertible, injective, surjective, one-to-one, onto,
Cayley graphs, Hedlund, Block representation, Lattice-gas automaton,
Reversible Cellular Automata.
Introduction
Cellular Automata (CA) consist in an array of identical cells, each of which
may take one in a finite number of possible states. The whole array evolves
in discrete time steps, and this global evolution is required to have a number
of physics-like symmetries: shift-invariance (it acts everywhere the same) and
causality (information cannot be transmitted faster than some fixed number of
cells per time step). CA are widely used to model spatially distributed computa-
tion (self-replicating machines, synchronization problems.. . ), as well as a great
variety of multi-agents phenomena (traffic jams, demographics.. . ). But their
origin lies in Physics, where they are commonly used to model waves or parti-
cles. In this context, it was natural to consider a further physics-like symmetry,
namely reversibility. The study of Reversible CA (RCA) was further motivated
by the promises of lower energy consumption in reversible computation. RCA
have turned out to have a beautiful mathematical theory, which relies on topol-
ogy or algebraic results to prove that 1/ the inverse of a CA is a CA [12], and
2/ that it admits a finite-depth reversible circuit decomposition [13].
CA model multi-agent systems, but under a fixed topology. There are many
situations, however, in which the notion of ‘who is next to whom’, also varies
in time (e.g. agents become physically connected, get to exchange contact de-
tails, move around, etc.). In the literature, several models (of physical systems,
self-replication, biochemical agents, economical agents, social networks. . . ) fea-
ture such neighbour-to-neighbour interactions with time-varying neighbourhood,
arXiv:1502.04368v1 [cs.DM] 15 Feb 2015
thereby generalizing CA for their specific sake. It is not until recently that CA
generalized to arbitrary, bounded-degree, time-varying graphs have been stud-
ied for their own sake, under the name of Causal Graph Dynamics (CGD). The
theoretical foundations of CGD have been laid in [1,3], including a first result
on reversibility (the inverse of a vertex-preserving CGD is a CGD). In this pa-
per, we improve this result, and prove that Reversible CGD admit a finite-depth
reversible circuit-like representation.
From a theoretical Computer Science viewpoint, the purpose of this paper is
therefore to generalize RCA theory to arbitrary, bounded-degree, time-varying
graphs. From a mathematical viewpoint, questions related to the bijectivity of
CA over certain classes graphs (more specifically, whether pre-injectivity implies
surjectivity for Cayley graphs generated by certain groups [10]) have received
some attention. This paper on the other hand provides a context in which to
study “bijectivity upon time-varying graphs”. This raises novel questions: Now
that the systems that support the information (i.e. the vertices) may be cre-
ated and deleted, does the bijectivity condition become vacuous? Or is it the
case that bijectivity will necessarily rigidify space (i.e. force the conservation of
each vertex)? We answer this question. From a theoretical physics viewpoint,
the question whether the reversibility of small scale physics (quantum mechan-
ics, micro-mechanical), can be reconciled with the time-varying topology of large
scale physics (relativity), is a topic of debate and constant investigation. This pa-
per provides a toy, discrete, classical model where reversibility and time-varying
topology coexist and interact.
1 Graphs
Basically, generalized Cayley graphs are the usual, connected, undirected, bounded-
degree graphs, but with five added twists:
Edges are between ports of vertices, rather than vertices themselves, so that
each vertex can distinguish its different neighbours, via the port that con-
nects to it.
There is a privileged vertex playing the role of an origin, so that any vertex
can be referred to relative to the origin, via a sequence of ports that lead to
it.
The graphs are considered modulo isomorphism, so that only the relative
position of the vertices can matter.
The vertices and edges are given labels taken in finite sets, so that they may
carry an internal state just like the cells of a Cellular Automaton.
The labelling functions are partial, so that we may express our partial knowl-
edge about part of a graph. For instance it is common that a local function
may yield a vertex, its internal state, its neighbours, and yet have no opinion
about the internal state of those neighbours.
Notations. Let πbe a finite set, Π=π2be its square, and V=P(Π) the set
of languages over the alphabet Π. The operator ‘.’ represents the concatenation
of words and εthe empty word, as usual.
The vertices of the graphs (see Figure 1(a)) we consider in this paper are uniquely
identified by a name uin V. Vertices may also be labelled with a state σ(u) in
Σa finite set. Each vertex has ports in the finite set π. A vertex and its port are
written u:a.
An edge is an unordered pair {u:a, v :b}. Such an edge connects vertices uand
v; we shall consider connected graphs only. The port of a vertex can only appear
in one edge, so that the degree of the graphs is always bounded by |π|. Edges
may also be labelled with a state δ({u:a, v :b}) in a finite set.
Formalization. Definitions 1 to 4 are as in [1]. The first two are reminiscent of
the many papers seeking to generalize Cellular Automata to arbitrary, bounded
degree, fixed graphs [17,7,11,10,9,22,15,21,20,6,16,8,18,19]. They are illustrated
by Figure 1(a).
Definition 1 (Graph). Agraph Gis given by
An at most countable subset V(G)of V, whose elements are called vertices.
A finite set π, whose elements are called ports.
A set E(G)of non-intersecting two-element subsets of V(G)×π, whose
elements are called edges. In other words an edge eis of the form {u:a, v :b},
and e, e0E(G), e e06=∅ ⇒ e=e0.
The graphs are all assumed to be connected, i.e. for any two u, v V(G), there
exists v1, . . . , vn1V(G)such that for all i∈ {0. . . n1}, one has {vi:ai, vi+1 :
bi} ∈ E(G)with v0=uand vn=v.
Definition 2 (Labelled graph). A labelled graph is a triple (G, σ, δ), also
denoted simply Gwhen it is unambiguous, where Gis a graph, and σand δ
respectively label the vertices and the edges of G:
σis a partial function from V(G)to Σ;
δis a partial function from E(G)to .
The set of all graphs with ports πis written Gπ. The set of all labelled graphs
with states Σ, ∆ and ports πis written GΣ,∆,π. To ease notations, we sometimes
write vGfor vV(G).
We now want to single out a vertex. The following definition is illustrated by
Figure 1(b).
Definition 3 (Pointed graph). Apointed (labelled) graph is a pair (G, p)
with pG. The set of all pointed graphs with ports πis written Pπ. The set of
all pointed labelled graphs with states Σ , ∆ and ports πis written PΣ,∆,π.
Definition 4 (Isomorphism). An isomorphism Ris a function from Gπto Gπ
which is specified by a bijection R(.)from Vto V. The image of a graph Gunder
the isomorphism Ris a graph RG whose set of vertices is R(V(G)), and whose
set of edges is {{R(u) : a, R(v) : b}|{u:a, v :b} ∈ E(G)}. Similarly, the image
of a pointed graph P= (G, p)is the pointed graph RP = (RG, R(p)). When P
and Qare isomorphic we write PQ, defining an equivalence relation on the
set of pointed graphs. The definition extends to pointed labelled graphs.
In the particular graphs we are considering, the vertices can be uniquely distin-
guished by the paths that lead to them starting from the pointer vertex. Hence,
we might just as well forget about vertex names. The following definition is
illustrated by Figure 1(c).
Definition 5 (Generalized Cayley graphs). Let Pbe a pointed (labelled)
graph (G, p). The Generalized Cayley graph Xis the equivalence class of Pwith
respect to the equivalence relation . The set of Generalized Cayley graphs with
ports πis written Xπ. The set of labelled Generalized Cayley graph with states
Σ, ∆ and ports πis written XΣ,∆,π .
These pointed graph modulo will constitute the set of configurations of the gen-
eralized Cellular Automata that we will consider in this paper.
u v
:a:b
(a)
u v
:a:b
(b)
:a:b
(c)
Fig. 1. The different types of graphs. (a) A graph. (b) A pointed graph.
(c) A pointed graph modulo or “Generalized Cayley graph”. These can also
be described as a language {ε, ab, ab.ba, . . .}and an equivalence relation with
equivalence classes corresponding to vertices: ˜ε={ε, ab.ba, . . .}and ˜
ab =
{ab, ab.ba.ab, . . .}
Paths and vertices. We mentioned that vertices can be uniquely distinguished
by the paths that lead to them starting from the pointer vertex. Let us make
this more precise.
Definition 6 (Path). Given a Generalized Cayley graph X, we say that αis
a path of Xif αis a sequence of ports aibisuch that, starting from the pointer,
it is possible to travel in the graph according to this sequence. More formally,
αis a path if there exists (G, p)Xand v1, . . . , vnV(G)such that for all
i∈ {0. . . n 1}, one has {vi:ai, vi+1 :bi} ∈ E(G), with v0=pand αi=aibi.
Notice that the existence of a path does not depend on the choice of (G, p)X.
Definition 7 (Equivalence of paths). Given a pointed graph modulo X, we
define the equivalence of paths relation Xsuch that for all paths u, u0of X,
uXu0if, starting from the pointer, uand u0lead to the same vertex of X. More
formally, uXu0if there exist (G, p)Xand v1, . . . , v|u|, v0
1, . . . , v0
|u0|V(G)
such that for all i∈ {0. . . |u| − 1},i0∈ {0. . . |u0| − 1}, one has {vi:ai, vi+1 :bi} ∈
E(G),{v0
i0:a0
i0, v0
i0+1 :b0
i0} ∈ E(G), with v0=p,v0
0=p,ui=aibi,u0
i0=a0
i0b0
i0
and v|u|=v|u0|. We write ˜ufor the equivalence class of umodulo X.
The vertices of a Generalized Cayley graph Xcan be designated by
˜uthe set of all paths leading to this vertex starting from the pointer of X
(in this convention the pointer is ˜ε),
or more directly by uan element of an equivalence class ˜u, i.e. a particular
path leading to this vertex starting from the pointer of X(in this convention
the pointer is ε).
In fact, given a Generalized Cayley graph we can reconstruct a pointed graph
(G(X),˜ε) according to these vertex naming conventions:
Definition 8 (Associated (pointed) graph). Let Xbe some Generalized
Cayley graph. Let P(X)be the pointed graph (G(X),˜ε), with G(X)such that:
The set of vertices V(G(X)) is the set of equivalence classes modulo Xof
the paths of X;
The edge {˜u:a, ˜v:b}is in E(G(X)) if and only if u.ab is a path of Xand
u.ab Xv, for all u˜uand v˜v.
We define the associated graph to be G(X). We define the associated pointed
graph to be P(X).
For convenience from now on ˜uand uwill no longer be distinguished. The latter
notation will be given the meaning of the former. I.e. we shall speak of a “vertex”
uin V(X) (or simply uX).
2 Operations
For a generalized Cayley graph (G, p) non-modulo (see [1] for details):
the neighbours of radius rare just those vertices which can be reached in r
steps starting from the pointer p;
the disk of radius r, written Gr
p, is the subgraph induced by the neighbours
of radius r+ 1, with labellings restricted to the neighbours of radius rand
the edges between them, and pointed at p.
Notice that the vertices of Gr
pcontinue to have the same names as they used
to have in G. For generalized Cayley graphs, on the other hand, the analogous
operation is:
Definition 9 (Disk). Let XXΣ,∆,π be a generalized Cayley graph and (G, ε)
its associated pointed graph. Let Xrbe f
Gr
ε. The generalized Cayley graph Xr
XΣ,∆,π is referred to as the disk of radius rof X. The set of disks of radius r
with states Σ, ∆ and ports πis written Xr
Σ,∆,π .
Definition 10 (Size). Let XXΣ,∆,π be a generalized Cayley graph. We say
that a vertex uXhas size less or equal to r+ 1, and write |u| ≤ r+ 1, if
uXr. We denote V(Xr
π) = SXXr
πV(X).
It will help to have a notation for the graph where vertices are named relatively
to some other pointer vertex u.
eε
eaa
ebb
gbb.ac
eca
eda
gda.cb
:a
:a
:b
:b
:a
:c
:c
:a
:d
:a
:b
:a
:c
:b
:c
:beε
eaa
ebb
eca
eda
:a
:a
:b
:b
:c
:a
:d
:a
XX0
Fig. 2. A generalized Cayley graph and its disk of radius 0.Notice that the
equivalence classes describing vertices in X0are strict subsets of those in X,
eventhough their shortest representative is the same. For instance the path ca.cb
is in ˜
da in Xbut is not a path in X0, and thus does not belong to ˜
da in X0.
Definition 11 (Shift). Let XXΣ,∆,π be a generalized Cayley graph and
(G, ε)its associated pointed graph. Consider uXor Xrfor some r, and con-
sider the pointed graph (G, u), which is the same as (G, ε)but with a different
pointer. Let Xube
^
(G, u). The generalized Cayley graph Xuis referred to as X
shifted by u.
The composition of a shift, and then a restriction, applied on X, will simply be
written Xr
u. Whilst this is the analogous operation to Gr
uover pointed graphs
non-modulo, notice that the shift-by-ucompletely changes the names of the
vertices of Xr
u. As the naming has become relative to u, the disk Xr
uholds no
information about its prior location u.
We may also want to designate a vertex vby those paths that lead to the
vertex urelative to ε, followed by those paths that lead to vrelative to u. The
following definition of concatenation coincides with the one that is induced by
the concatenation of words belonging to the classes uand v:
Definition 12 (Concatenation). Let XXπbe a generalized Cayley graph
and (G, ε)its associated pointed graph. Consider uXand vXuor Xr
ufor
some r. Let (G0, ε)be the associated pointed graph of (Xu)v,Rbe an isomorphism
such that G0=RG, and u.v be R1(ε). The vertex u.v Xis referred to as u
concatenated with v.
According to Definition 11, G0and Gare isomorphic. Moreover, the restriction
of R1to V(G0) is uniquely determined; hence the definition is sound.
It also helps to have a notation for the paths to εrelative to u.
Definition 13 (Inverse). Let XXπbe a generalized Cayley graph and (G, ε)
its associated pointed graph. Consider uX. Let (G0, ε)be the associated pointed
graph of Xu,Rbe an isomorphism such that G0=RG, and ube R(ε). The vertex
uXuis referred to as the inverse of u.
Notice the following easy facts: (Xu)v=Xu.v,u.u =ε. Notice also that the
isomorphism Rsuch that G(Xu) = RG(X) maps vto u.v.
3 Causality
This notion of causality extends the known mathematical definition of Cellular
Automata over grids and Cayley graphs. The extension will be a strict one for
two reasons: not only the graphs become arbitrary, but they can also vary in
time.
The main difficulty we encountered when elaborating an axiomatic definition
of causality from XΣ,∆,π to XΣ,∆,π , was the need to establish a correspondence
between the vertices of a generalized Cayley graph X, and those of its image
F(X). Indeed, on the one hand it is important to know that a given uXhas
become u0F(X), e.g. in order to express shift-invariance F(Xu) = F(X)u0.
But on the other hand since u0is named relative to ε, its determination requires
a global knowledge of X.
The following analogy provides a useful way of tackling this issue. Say that
we were able to place a white stone on the vertex uXthat we wish to
follow across evolution F. Later, by observing that the white stone is found at
u0F(X), we would be able to conclude that uhas become u0. This way of
grasping the correspondence between an image vertex and its antecedent vertex
is a local, operational notion of an observer moving across the dynamics.
Definition 14 (Dynamics). A dynamics (F, R)is given by
a function F:XΣ,∆,π XΣ,∆,π;
a map R, with R:X7→ RXand RX:V(X)V(F(X)).
For all X, the function RXcan be pointwise extended to sets, i.e. RX:P(V(X))
P(V(F(X))) maps Sto RX(S) = {RX(u)|uS}.
The intuition is that RXindicates which vertices {u0, v0, . . .}=RX({u, v, . . .})
V(F(X)) will end up being marked as a consequence of {u, v, ... X} ⊆ V(X)
being marked. Now, clearly, the set {(X, S)|XXΣ ,∆,π, S V(X)}is isomor-
phic to XΣ0,∆,π with Σ0=Σ× {0,1}. Hence, we can define the function F0that
maps (X, S)
=X0XΣ0,∆,π to (F(X), RX(S))
=F0(X0)XΣ0,∆,π, and think
of a dynamics as just this function F0:XΣ0,∆,π XΣ0,∆,π. This alternative
formalism will turn out to be very useful.
Definition 15 (Shift-invariance). A dynamics (F, R)is said to be shift-
invariant if for every Xand uX,vXu,
F(Xu) = F(X)RX(u)
RX(u.v) = RX(u).RXu(v).
The second condition expresses the shift-invariance of R. Notice that RX(ε) =
RX(ε).RX(ε); hence RX(ε) = ε.
In the F0:XΣ0,∆,π XΣ0,∆,π formalism, the two above conditions are equivalent
to just one: F0(X0
u) = F0(X0)RX(u).
Definition 16 (Uniform continuity). A dynamics (F, R)is said to be con-
tinuous if for all m, there exists nsuch that for all X, X 0,X0n=Xnimplies
both
F(X0)m=F(X)m.
dom Rm
X0V(X0n), dom Rm
XV(Xn)and Rm
X0=Rm
X.
where Rm
Xdenotes the partial map obtained as the restriction of RXto the
codomain F(X)m, using the natural inclusion of F(X)minto F(X).
In the F0:XΣ0,∆,π XΣ0,∆,π formalism, the two above conditions are equivalent
to just one: F0uniformly continuous.
We need one third, last condition:
Definition 17 (Boundedness). A dynamics (F, R)from XΣ,∆,π to XΣ,∆,π is
said to be bounded if there exists a bound bsuch that for all X, for all w0F(X),
there exist u0im RXand v0F(X)b
u0such that w0=u0.v0.
The following is the definition of causality:
Definition 18 (Causal dynamics). A dynamics is causal if it is shift-invariant,
uniformly continuous and bounded.
Actually, uniform continuity can be weakened into continuity, due to the com-
pactness of XΣ,∆,π , see [2].
An example of causal dynamics is the inflating grid dynamics illustrated in
Figure 3. In the inflating grid dynamics each vertex gives birth to four distinct
vertices, such that the structure of the initial graph is preserved, but inflated.
The graph has maximal degree 4, and the set of ports is π={a, b, c, d}, vertices
and edges are unlabelled.
4 Locality
Causal Graph Dynamics change the entire graph in one go. The word causal there
refers to the fact that information does not propagate too fast. Local operations,
on the other hand, act just in one bounded region of the graph, leaving the rest
unchanged. We introduce the following locality definition:
Definition 19 (Local dynamics). A dynamics (L, S)is r0-local if it is uni-
formly continuous and bounded, and there exists r0such that for all Xand
u0L(X)with |u0|> r0, there exists uXsuch that we have both:
L(X)0
u0=X0
u,
• ∀vX0
u, SX(u.v) = u0.v.
Lemma 1 (Bounded inflation). If (L, S)is r0-local, then for all sthere exists
s0such that for all Xand vX, if |v| ≤ s, then |SX(v)| ≤ s0.
7→
Fig. 3. The inflating grid dynamics. Each vertex splits into 4 vertices. The
structure of the grid is preserved. Ports are omitted here.
Proof. Suppose the contrary: X(s0) has some |v(s0)| ≤ ssuch that |SX(v)|> s0.
Since XΣ,∆,π is compact [2], X(s0) admits a subsequence which converges to some
limit X, in the sense that X(s0
k)k=Xk. For this particular X, for any s0, there is
some |v(s0)| ≤ ssuch that |SX(v)|> s0. This is because we can choose kso that
s0
ks0and ksuperior to the radius needed to determine L(X)s0
k=L(X(s0
k))s0
k,
so that |SX(v)|=|SXk(v)|=|SX(sk)k(v)|=|SX(sk)(v)|> s0
ks0. Thus, there
exists a point of vXhas |SX(v)|>, which is a contradiction.
Lemma 2. If (L, S)is r0-local, for all t, for all u0L(X)with |u0|> r0+t+1,
there exists uXwith SX(u) = u0such that we have:
L(X)t
u0=Xt
u,
• ∀vXt
u, SX(u.v) = u0.v.
Proof. Take such a u0and consider usuch that u0=SX(u).
[First ] Since |u0|> r0+t+ 1, we have that for all vL(X)t
u0,|u0.v|> r0.
Hence, by r0-locality of L, there exists xXsuch that SX(x) = u0.v and such
that L(X)0
u0.v =X0
x, i.e. the vertex vin L(X)t
u0, in terms of its internal states
and edges, is the same as the vertex xin X. Now, say there exists |z|= 1 such
that w=v.z L(X)t
u0, i.e. there is an edge between vand v.z in L(X)t
u0. Again
since |SX(x)|> r0, the r0-locality yields u0.v.z =SX(x).z =SX(x.z), i.e. the
edge between vand v.z v in L(X)t
u0is the same as that between xand x.z in
X. Consider v1. . . vk=vwith ktand |vi|= 1. A similar argument starting
from u0and following these edges shows that xis at distance tof uin X, and
thus x.z is at distance t+ 1 of uin X. So the vertices x,x.v and their edge do
appear in Xt
u.
[Second ] Again take wXt
u=L(X)t
u0. Consider w1. . . wk=wwith kt+ 1
and |wi|= 1. Since |u0|> r0+t+ 1 > r0, the r0-locality applies and yields
SX(u.w1) = SX(u).w1=u0.w1. Similarly, since |u0.w1. . . wi|> r0+t+ 1 i>r0,
the r0-locality applies and yields SX(u.w1. . . wi.wi+1) = SX(u.w1. . . wi).wi+1 =
u0.w1. . . wi.wi+1. Eventually SX(u.w) = u0.w.
Local operations do not need to act just over the region surrounding the
origin. We may also shift them to act over the region surrounding some vertex
u.
Definition 20 (Shifted dynamics). Consider a dynamics (L, S)and some
uΠ. We define Luto be the map X7→ (L(Xu))SXu(u)if uX, and the
identity otherwise. We define Su,X to be the map v7→ SXu(u).SXu(u.v)if uX,
and the identity otherwise. We say that (Lu, Su,)is (L, S) shifted at u.
We may wish to apply a series of local operations at different positions ui, i.e.
a circuit. However, applying a local operation may change the graph and hence
vertex names, hence some care must be taken.
Definition 21 (Product). Consider a local dynamics (L, S)and Xa gen-
eralized Cayley graph in its domain we define the product QL(X)as the limit
when rgoes to infinity of:
Y(r) = Y
i[1,...,|V(Xr)|]
Lu0
i(X)
where {u1, u2, ...}=V(Xr),u0
1=u1,u0
2=Su0
1,X (u2),u0
3=Su0
2,Lu0
1(X)(u3),...
5 Invertibility and almost-vertex-preservingness
The following definition imposes invertibity in a natural fashion.
Definition 22 (Invertible dynamics). A dynamics (F, R)is said to be in-
vertible if Fis a bijection over XΣ,∆,π.
Recall that, in general, CGD are allowed to evolve the graph, not only by
changing internal states and edges, but also by creating or deleting vertices.
Since invertibility imposes information-conservation, one may wonder whether
invertible CGD are still allowed to create or delete vertices. Indeed, they are, as
show by Figure 5. One notices, however, that the RHS of this example features
shift-equivalent vertices.
Definition 23 (Shift-equivalent vertices). Let XXΣ,∆,π and let u, v
V(X). We say that uand vare shift-equivalent if Xu=Xv. This equivalence
relation is denoted uv. A graph is called asymmetric if it has only trivial (i.e.
of size one) shift-equivalence classes.
One can show that the shift-equivalence classes of generalized Cayley graph
must always have the same size. Intuitively, if you consider two shift-equivalent
vertices, you can travel along a given path starting from either of the them, this
will give you two distinct equivalent vertices.
X
ε
u
v
u
Xu
u
ε
u.v L
L(Xu)
SXu(u)
ε
SXu(u.v)
Lu
L(Xu)SXu(u.v)
ε
SXu(u)
SXu(u.v)
Fig. 4. Shift-invarience of a shifted dynamics Lu. In the bottom graph Lu(X),
former vertex vhas name SXu(u).SXu(u.v).
Fig. 5. The turtle dynamics has the two above generalized Cayley graphs to
oscillate between one another. The two vertices of the RHS are shift-equivalent,
pointing the graph upon one or the other does not change the graph.
Lemma 3 (Shift-equivalence classes isometry). Let XXΣ,∆,π be a graph.
If C1V(X)and C2V(X)are two shift-equivalence classes of X, then
|C1|=|C2|.
Proof. Consider two equivalent and distinct vertices uand vin X. Consider a
path w. The vertices u.w and v.w are distinct and equivalent. More generally,
if we have nequivalent distinct vertices v1, ..., vn, any vertex u=v1.w will be
equivalent to v2.w, ..., vn.w and distinct from all of them, hence the equivalence
classes are all of the same size.
Shift-symmetry is fragile, however, and can be destroyed by adding a few
vertices to it:
Definition 24 (Primal extension). Given a finite graph XXΣ,∆,π where
|π|>1such that Xhas kshift-equivalence classes of size nwith k, n 6= 1,
consider the following transformation:
If Xhas a free port: connect pk.n new vertices in a line to this free port,
where pis the second smallest prime number greater than k.n
If Xhas no free port: Xhas at least one cycle. Remove an edge of this cycle,
and do the same construction as above.
We denote as 2Xthis new graph.
Lemma 4 (Properties of primal extensions). Any primal extension 2Xis
asymmetric and |2X| − |X|=o(|X|)when |X|tends toward infinity, where |X|
is the size of the graph X.
Proof. As 2Xhas a prime number of vertices, by lemma 3, its has either one
single equivalence class of maximal size or only trivial equivalence classes. As the
2.operation adds at least 2 vertices and that these vertices have different degree
(1 for the last vertex on the line, and 2 for its only neighbor), 2Xcontains at
least two non equivalent vertices, hence the first result. Moreover, according to
the prime number theorem, the quantity |2X|−|X|is of the order of log(|X|),
hence the second result.
This primal extension construction is the key ingredient to prove that invertible
CGD cannot exploit symmetries of a graph to create or delete vertices. Indeed, by
considering a graph whose size is changed through the application of an invertible
CGD and considering its primal extension, we can contradict the continuity the
of the invertible CGD.
Lemma 5 (Invertible preserves shift-equivalence classes). Let (F, R)be
a shift-invariant dynamics over Xπ, such that Fis a bijection. Then for all graph
X, for any u, v in Xwe have uvis equivalent to RX(u)RX(v).
Proof. uvexpresses Xu=Xv, which by bijectivity of Fis equivalent to
F(Xu) = F(Xv) and hence F(X)RX(u)=F(X)RX(v). This in turn is expressed
by RX(u)RX(v).
Using this fact and the primal extension construction, one can show that the
cases of node creation and deletion in invertible CGD are all of finitary nature,
i.e. they can no longer happen for large enough graphs. Intuitively, this is be-
cause according to the previous Proposition, creation or deletion of vertices must
respect to the shift-symmetries of the graph. But these are global properties, and
thus cannot be exploited by an invertible CGD.
Lemma 6 ((Finitely)-almost-vertex-preserving). Let (F, R)be an invert-
ible causal graph dynamics over XΣ,∆,π , such that there exists a bound p, such
that for all finite graph X, if |X|> p then RXis bijective. Then Xinfinite
implies RXbijective.
Proof. [RXinjective ]. By contradiction. Take Xinfinite such that there is
u6=vand RX(u) = RX(v). Without loss of generality we can take u=ε, i.e.
v6=εand RX(v) = ε. By continuity of R, there exists a radius r, which we can
take larger than |v|and p, such that RX=RXr. Then RXr(v) = RX(v) = ε, thus
RXris not injective in spite of Xrbeing finite and larger than p, a contradiction.
[RXsurjective ]. By contradiction. Take Xinfinite such that there is v0in
F(X) and v0/im RX. By boundedness and shift-invariance, we can assume
that there exists bsuch that |v0|< b. By continuity of R, there exists a radius
r, which we can take larger than p, such that the images of RXand RXrcoincide
over the disk of radius b. Then, v0/im RXimplies v0/im RXr, thus RXris
not surjective in spite of Xrbeing finite and larger than p, a contradiction.
Theorem 1 (Invertible implies almost-vertex-preserving). Let (F, R)
be a causal graph dynamics over XΣ,∆,π, such that Fis a bijection. Then there
exists a bound p, such that for all graph X, if |X|> p then RXis bijective.
Proof. Let us prove this result for all finite graphs. Then, using lemma 6, the
general result will follow.
By contradiction, let us assume that there exists a sequence of finite graphs
(X(n))nNsuch that |X(n)|diverges and such that for all n,RX(n)is not bijec-
tive. As this sequence is infinite, we have that one of the two following cases is
verified an infinite number of time:
RX(n)is not surjective,
RX(n)is not injective.
[RX(n)not surjective]. There exists a vertex v0/im RX(n). Without loss of
generality, we can assume that |v0|< b where bis the bound from the bounded-
ness property of F. Consider the graph Y(n) = F1(2F(X(n))). Using uniform
continuity of F1and R, and the fact that |X(n)|is as big as we want, we
have that there exists an index nand a radius rsuch that Y(n)r=X(n)rand
Rb
Y(n)r=Rb
X(n)r. As F(Y(n)) is asymmetric by construction, v0im Rb
Y(n)r
which contradicts v0/im RX(n).
[RX(n)not injective]. There exists two vertices u, v X(n) such that RX(n)(u) =
RX(n)(v) and u6=v. Without loss of generality, we can assume that u=εas
Fis shift-invariant. According to lemma 5, we have that εv. Moreover, us-
ing the uniform continuity of R, we have that, as RX(n)(v) = RX(n)(ε) = ε,
there exists a radius l, which does not depend on n, such that |v|< l . Let
us consider the graph 2X(n). In this graph, εand vare not shift-equivalent
and thus, R2X(n)(ε)6=R2X(n)(v) . By continuity of R, we have that there
exists a radius r > l such that R0
2X(n)r=R0
X(n)rfor a large enough n, hence
R0
2X(n)r(v) = R0
X(n)r(v) = ε, which contradicts R2X(n)(ε)6=R2X(n)(v).
In [3] it is shown that the inverses of vertex-preserving invertible CGD are
also invertible.
Definition 25 (Vertex-preserving invertible). A shift-invariant dynamics
(F, R)is vertex-preserving invertible if Fis a bijection and for all Xwe have
that RXis a bijection.
More precisely, we know from [3] that if (F, R) is a vertex-preserving invert-
ible shift-invariant dynamics, then (F1, S) is a shift-invariant dynamics, with
SY= (RF1(Y))1. We also know that if (F, R) is a vertex-preserving invert-
ible causal dynamics, then it is reversible, i.e. (F1, S) is also a causal dynam-
ics, with SY= (RF1(Y))1. The above theorem now shows that the vertex-
preservingness assumption was almost without loss of generality. Thus, it is easy
to extend the result to any invertible CGD.
Corollary 1 (Invertible implies reversible). Let (F, R)be an invertible
causal graph dynamics over Xπ,∆,Σ . Then there exists a Ssuch that (F1, S)
is a causal graph dynamics, i.e. (F, R)is reversible.
Proof. We must construct S. For |F(X)|=|X|> p, we know that RXis bijec-
tive and we let SF(X)=R1
X. For |X| ≤ p, we proceed as follows.
We write ˜ufor the shift-equivalence class of u. For all v0F(X), we make
the arbitrary choice SF(X)(˜
v0) = v, where vis such that RX(v)v0. For
this X, we have enforced -compatibility. In order to enforce shift-invariance,
we must make consistent choices for SF(X)u0. This is obtained by demanding
that SF(X)u0(g
u0.v0) = u.v. Indeed, this accomplishes shift-invariance because
SF(X)u0(v0) = SF(X)u0(u0.u0.v0) = ε.v0=v0implying the equality: SF(X)(u0.v0) =
u.v =SF(X)(u0).SF(X)u0(v0). Moreover, SF(X)u0is itself shift-invariant because:
SF(X)u0.v0(w0) = SF(X)u0.v0(v0.v0.w0) = ε.w =wand SF(X)u0(v0) = vimply-
ing that SF(X)u0(v0.w0) = v.w =SF(X)u0(v0).SF(X)u0.v 0(w0) , and -compatible
because v0w0implies SF(X)u0(v0) = SF(X)u0(w0), and thus SF(X)u0(v0)
SF(X)u0(w0).
Continuity of the constructed Sis due to the continuity of Rand the finiteness
of p.
Shift-invariance of (F1, S) follows from -compatibility of Sand shift-invariance
of (F, R), because F1(F(X)0
u) = Xvwhere vis such that RX(v)u0, hence
F1(F(X)0
u) = XSF(X)(u0).
Thus, invertible CGD are in fact Reversible CGD (RCGD).
6 Block representation
A famous result on RCA [13], is that these admit a finite-depth, reversible circuit
form, with gates acting only locally. The result carries through to Quantum
CA [5], whose proof technique inspired the following [4]. First, we show that
conjugating a local operation with an RCGD still yields a local operation.
Proposition 1. If (F, R)is an RCGD and (L, S)is local, then (L0, T)is
local, with L0=F1LFand TX(u)=(RF1(L(F(X)))1(SF(X)(RX(u))).
Proof. Boundedness and uniform continuity by composition. Next, suppose: Lis
local, r0is such that for all X, Y if Xr0=Yr0then F1(X)0=F1(Y)0(given
by uniform continuity of F1), r2bFis such that for all X, Y if Xr2bF=Yr2bF
then F(X)2bF=F(Y)2bF(given by uniform continuity of F), bF1is the bound
given by the bounded inflation lemma applied on F1,bLis the bound given by
the boundedness of Land rLthe radius of locality of L. In the two following
points, we chose a radius r0as follow :
r0=bF1(rL+2+max(r0,2bF, r2bF))
Consider |u0|> r0.
[First ] Let us show that there exists uXsuch that L0(X)0
u0=X0
u. By
definition of F1, there exists wLF (X) such that RF1
LF (X)(w) = u0. By
bounded inflation of F1, we have |w|> rLand thus by locality of L, there
exists w0F(X) such that SF(X)(w0) = w. Finally by reversibility of F
there exists uXsuch that RF
X(u) = w, and thus u0=TX(u). Notice
that we have that |SF(X)RF
X(u)|> r0+rL+ 2. Using lemma 2 with t=
r0, we have: LF (X)r0
SF(x)RF
X(u)=F(X)r0
RF
X(u)=F(Xu)r0. By definition of r0,
F(Xu)r0=LF (X)r0
SF(x)RF
X(u)implies X0
u=F1(LF (X)r0
SF(x)RF
X(u))0, which
leads by shift-invariance of F1to X0
u=F1(LF (X)r0)0
RF1
LF (X)SF(x)RF
X(u)Hence
X0
u=L0(X)0
u0.
[Second ] Consider uas above and vX0
u.
TX(u.v) = RF1
LF (X)SF(x)(RF
X(u.v))
=RF1
LF (X)SF(x)(RF
X(u).RF
Xu(v)) using shift-invariance of F
=RF1
LF (X)SF(x)(RF
X(u)).RF
Xu(v) because |SF(X)RF
X(u)|> rL+ 2bF+ 2
=RF1
LF (X)(SF(x)(RF
X(u))).RF1
LF (X)SF(X)(RF
X(u)) (RF
Xu(v)) using shift-invariance of F1
=TX(u).RF1
LF (X)SF(X)(RF
X(u)) (RF
Xu(v))
We will now show that: v=RF1
LF (X)SF(X)(RF
X(u)) (RF
Xu(v)). Since |RF
Xu(v)|<2bF
by bounded inflation of F, it is enough to show:
RF1
LF (X)SF(X)(RF
X(u))
2bF(RF
Xu(v)) = v.
By definition of r2bF, we have that: if Xr2bF=Yr2bFthen RF1(X)2bF=
RF1(Y)2bF. Let us show that LF (X)r2bF
S2bF
F(X)(RF
X(u)) =F(Xu)r2bF. By apply-
ing lemma 2 with t=r2bF,LF (X)r2bF
S2bF
F(X)(RF
X(u)) =F(X)r2bF
RF
X(u)which, by shift-
invariance of F, is equal to F(Xu)r2bF. As a consequence,
RF1
LF (X)SF(X)(RF
X(u))
2bF(RF
Xu(v)) = RF1
F(Xu)
2bF(RF
Xu(v)) = RF1
F(Xu)(RF
Xu(v)) = v
by definition of RF1
Second, we give ourselves a little more space so as to mark which parts of
the graph have been update, or not.
Definition 26 (Marked generalized Cayley graphs). Consider the set of
generalized Cayley graphs XΣ,π with labels in Σ, and ports in π. Let Σ0=Σ×
{0,1}and π0=π×{0,1}. We define the set of marked generalized Cayley graphs
XΣ00to be the subset of XΣ00such that:
for all uXwith σX(u) = (x, a)and {u: (i, b), v : (j, c)} ∈ X, we have
a=c.
for all vXwith {u: (i, b), v : (j, c)} ∈ Xand {u0: (i0, b0), v : (j, c0)} ∈ X,
we have u=u0.
Definition 27 (Mark operation). Given a a label in Σ0or of a port in π0,
we define the mark operation µ(.), as flipping the bit in the second component.
Given a graph Xin XΣ00, we define the mark operation, µ:XΣ00XΣ00
as follows:
σµ(X)(ε) = µ(σX(ε))
For all i, j π0,{ε:µ(i), ε :µ(j)} ∈ µ(X)if {ε:i, ε:j} X.
For all vXwith v6=εand i, j π0,{ε:i, v :j0} ∈ µ(X)if {ε:i, v :j} ∈ X.
and leaving the rest of the graph Xunchanged.
Notice that the set of marked graphs XΣ00is but the subset of XΣ00obtained
by as closure of µ, and shifts, upon Xπ×{0}×{0}.
Remark: The set XΣ00is a compact subset of Xπ00,∆ .
It turns out that any RCGD admits an extension that allows for these marks.
Definition 28 (Reversible extension). Let F:Xπ,Σ,∆ Xπ,Σ,∆ be an
RCGD. We say that F0:XΣ00XΣ00is a reversible extension of Fif
F0is an RCGD, and:
For all XXΣ×{0}×{0},F0(X) = F(X).
For all XXΣ×{1}×{1},,F0(X) = X.
For all |X| ≤ pand X /XΣ×{0}×{0},F0(X) = X, where pis that of
Theorem 1.
Proposition 2 (Reversible extension). Suppose F:Xπ,Σ,∆ XΣ is an
RCGD. Then it admits a reversible extension F0:XΣ00XΣ00.
Proof. See appendix A.
In order to obtain our circuit-like form for RCGD, we will proceed by re-
versible, local updates.
Definition 29 (Conjugate mark). Given a reversible extension F0:XΣ00
XΣ00, we define the conjugate mark K:XΣ00XΣ0 0to be the function:
K=F0−1µF0.
Notice that by Proposition 1, the local update blocks are local operations. More-
over, since they are defined as a composition of invertible dynamics, so they are.
In order to represent the whole of an RCGD, it suffices to apply these local
update blocks at every vertex.
Theorem 2 (Reversible localizability). Suppose F:Xπ,Σ ,∆ XΣ,π is an
RCGD. Then, there exists psuch that for all XXπ,Σ,∆, if |X|> p we have:
F(X)=(Yµ)(YK)(X)
where K=F0−1µF 0for F0:XΣ00XΣ00a reversible extension of F.
Proof. Take XXπ,Σ,∆. We have:
(Yµ)(YK)(X)=(Yµ)(YF0−1µF 0)(X)
= (Yµ)F0−1(Yµ)F0(X)
= (Yµ)(Yµ)F(X)
=F(X).
Notice that the cases |X|< p are finite and Fis bijective, thus it just permutes
those cases. Thus, this theorem generalizes the block decomposition of reversible
cellular automata, which represents any reversible cellular automata as a circuit
of finite depth of local permutations. Here, the mark µand its conjugate Kare
the local permutations. The circuit is again of finite depth, a vertex uwill be
attained by all those Kthat act over Xr0
u, where r0is the locality radius of K.
Therefore, the depth is less than |π|r0.
7 Conclusion
Summary of results. We have studied Reversible Causal Graph Dynamics, thereby
extending Reversible Cellular Automata results to time-varying, Generalized
Cayley graphs. Generalized Cayley Graphs are arbitrary bounded-degree net-
works, with a pointed vertex serving as the origin, and modulo renaming of
vertices. Some of these graphs have shift-equivalent vertices. We have shown
that if a Causal Graph Dynamics (CGD) is invertible, then it preserves shift-
equivalence classes. This in turn entails almost-vertex-preservingness, i.e. the
conservation of each vertex but for big enough graphs. Combining this with
earlier results on vertex-preserving invertible CGD, we have shown that the in-
verse of a CGD is a CGD. Next, we have investigated whether these Reversible
CGD can be implemented by small, invertible, local operations. A local opera-
tion is one which acts upon a subdisk of the graph, leaving the rest unchanged.
Our result shows that any Reversible CGD can be simulated by another, which
proceeds via a bounded-radius, invertible, commutative updates, followed by a
0-radius, invertible, marking operation.
Future work. These are generalization of theoretical Computer Science results
[12,13]. From a mathematical point of view, however, they show that invertibility
is a very strong constraint on space-varying dynamics: beyond some finitary
cases, information conservation implies conservation of the systems that support
this information. Still, this cannot forbid that some ‘dark matter’ which was
there at all times, could now be made visible. We plan to follow this idea in a
subsequent work. We also wish to explore the quantum regime of these models,
as similar results where given for Quantum Cellular Automata over fixed graphs
[5,4], whose methods have inspired the block representation theorem of this
paper. Such results would be of interest to theoretical physics, in the sense of
discrete time versions of [14].
Acknowledgements
This work has been funded by the ANR-12-BS02-007-01 TARMAC grant, the
ANR-10-JCJC-0208 CausaQ grant, and the John Templeton Foundation, grant
ID 15619. The authors acknowledge enlightening discussions with Bruno Mar-
tin and Emmanuel Jeandel. This work has been partially done when PA was
delegated at Inria Nancy Grand Est, in the project team Carte.
References
1. P. Arrighi and G. Dowek. Causal graph dynamics. In Proceedings of ICALP 2012,
Warwick, July 2012, LNCS, volume 7392, pages 54–66, 2012.
2. P. Arrighi and S. Martiel. Generalized Cayley graphs and cellular automata
over them. In Proceedings of GCM 2012, Bremen, September 2012. Pre-print
arXiv:1212.0027, pages 129–143, 2012.
3. P. Arrighi, S. Martiel, and V. Nesme. Generalized Cayley graphs and cellular
automata over them. submitted (long version). Pre-print arXiv:1212.0027, 2013.
4. P. Arrighi and V. Nesme. A simple block representation of Reversible Cellular
Automata with time-symmetry. In 17th International Workshop on Cellular Au-
tomata and Discrete Complex Systems, (AUTOMATA 2011), Santiago de Chile,
November 2011., 2011.
5. P. Arrighi, V. Nesme, and R. Werner. Unitarity plus causality implies localizability.
J. of Computer and Systems Sciences, 77:372–378, 2010. QIP 2010 (long talk).
6. P. Boehm, H.R. Fonio, and A. Habel. Amalgamation of graph transformations:
a synchronization mechanism. Journal of Computer and System Sciences, 34(2-
3):377–408, 1987.
7. B. Derbel, M. Mosbah, and S. Gruner. Mobile agents implementing local compu-
tations in graphs. Graph Transformations, pages 99–114, 2008.
8. H. Ehrig and M. Lowe. Parallel and distributed derivations in the single-pushout
approach. Theoretical Computer Science, 109(1-2):123–143, 1993.
9. Tullio Ceccherini-Silberstein; Francesca Fiorenzi and Fabio Scarabotti. The Garden
of Eden Theorem for cellular automata and for symbolic dynamical systems. In
Random walks and geometry. Proceedings of a workshop at the Erwin Schr¨odinger
Institute, Vienna, June 18 – July 13, 2001. In collaboration with Klaus Schmidt
and Wolfgang Woess. Collected papers., pages 73–108. Berlin: de Gruyter, 2004.
10. M. Gromov. Endomorphisms of symbolic algebraic varieties. Journal of the Euro-
pean Mathematical Society, 1(2):109–197, April 1999.
11. Stefan Gruner. Mobile agent systems and cellular automata. Autonomous Agents
and Multi-Agent Systems, 20:198–233, 2010. 10.1007/s10458-009-9090-0.
12. G. A. Hedlund. Endomorphisms and automorphisms of the shift dynamical system.
Math. Systems Theory, 3:320–375, 1969.
13. J. Kari. Representation of reversible cellular automata with block permutations.
Theory of Computing Systems, 29(1):47–61, 1996.
14. T. Konopka, F. Markopoulou, and L. Smolin. Quantum graphity. Arxiv preprint
hep-th/0611197, 2006.
15. H.J. Kreowski and S. Kuske. Autonomous units and their semantics-the parallel
case. Recent Trends in Algebraic Development Techniques, pages 56–73, 2007.
16. M. L¨owe. Algebraic approach to single-pushout graph transformation. Theoretical
Computer Science, 109(1-2):181–224, 1993.
17. C. Papazian and E. Remila. Hyperbolic recognition by graph automata. In Au-
tomata, languages and programming: 29th international colloquium, ICALP 2002,
alaga, Spain, July 8-13, 2002: proceedings, volume 2380, page 330. Springer Ver-
lag, 2002.
18. G. Taentzer. Parallel and distributed graph transformation: Formal description and
application to communication-based systems. PhD thesis, Technische Universitat
Berlin, 1996.
19. G. Taentzer. Parallel high-level replacement systems. Theoretical computer science,
186(1-2):43–81, 1997.
20. K. Tomita, S. Murata, A. Kamimura, and H. Kurokawa. Self-description for con-
struction and execution in graph rewriting automata. Advances in Artificial Life,
pages 705–715, 2005.
21. Kohji Tomita, Haruhisa Kurokawa, and Satoshi Murata. Graph automata: natural
expression of self-reproduction. Physica D: Nonlinear Phenomena, 171(4):197 –
210, 2002.
22. Kohji Tomita, Haruhisa Kurokawa, and Satoshi Murata. Graph-rewriting au-
tomata as a natural extension of cellular automata. In Thilo Gross and Hiroki
Sayama, editors, Adaptive Networks, volume 51 of Understanding Complex Sys-
tems, pages 291–309. Springer Berlin / Heidelberg, 2009.
A Proof of Proposition 3
The next two definitions are standard, see [6,16] and [1]. Basically, we need a
notion of union of graphs, and for this purpose we need a notion of consistency
between the operands of the union:
Definition 30 (Consistency). Let XXπbe a generalized Cayley graph. Let
Gbe a labelled graph (G, σ, δ), and G0be a labelled graph (G0, σ0, δ0), each one
having vertices that are pairwise disjoint subsets of V(X). The graphs are said
to be consistent if:
(i) xGx0G0xx06=∅ ⇒ x=x0,
(ii) x, y Gx0, y0G0a, a0, b, b0π({x:a, y :b} ∈ E(G)∧ {x0:a0, y0:b0} ∈
E(G0)x=x0a=a0)(b=b0y=y0),
(iii) x, y Gx0, y0G0a, b π x =x0δ({x:a, y :b}) = δ0({x0:a, y0:b})
when both are defined,
(iv) xGx0G0x=x0σ(x) = σ0(x0)when both are defined.
The consistency conditions aim at making sure that both graphs “do not dis-
agree”. Indeed: (iv) means that “if Gsays that vertex xhas label σ(x), G0should
either agree or have no label for x”; (iii) means that “if Gsays that edge ehas
label δ(e), G0should either agree or have no label for e”; (ii) means that “if G
says that starting from vertex xand following port aleads to yvia port b,G0
should either agree or have no edge on port x:a”.
Condition (i) is in the same spirit: it requires that Gand G0, if they have a vertex
in common, then they must fully agree on its name. Remember that vertices of
Gand G0are disjoint subsets of V(X). If one wishes to take the union of Gand
G0, one has to enforce that the vertex names will still be disjoint subsets of V(X).
Definition 31 (Union). Let XXπbe a generalized Cayley graph. Let G
be a labelled graph (G, σ, δ), and G0be a labelled graph (G0, σ0, δ0), each one
having vertices that are pairwise disjoint subsets of V(X).S. Whenever they are
consistent, their union is defined. The resulting graph GG0is the labelled graph
with vertices V(G)V(G0), edges E(G)E(G0), labels that are the union of the
labels of Gand G0.
Definition 32 (Upper and lower projections). Let Gbe a graph in G(XΣ00).
We define
G(resp.
G) the lower (resp. upper) projection of Gas the set of the
connected component obtained after removing all marked vertices (resp. all non-
marked vertices without used marked ports).
Lemma 7 (Characterization of connected components). Given Gin G(XΣ00),
the elements of the sets
Gand
Gare of the form u.Y with u˜
Gand Y
G(XΣ00).
Proposition 3 (Reversible extension). Suppose F:Xπ,Σ,∆ XΣ is an
RCGD. Then it admits a reversible extension F0:XΣ00XΣ00.
Proof. Let us construct such a reversible extension F0. Let pbe that of Theorem
1. For all |X| ≤ pand X /XΣ×{0}×{0}, we let F0(X) = X. The rest supposes
|X|> p.
Given L:XΣ,π XΣ,π , we define L?as the function GF. Now for all
XXΣ00, we define F0(X) as the equivalence class modulo isomorphism of
the following graph pointed on ε:
[
C
G
C
[
u.Y
G
u.F ?(˜
Y)
where G=G(X). Notice that if GGΣ×{0}×{0}then
Gis empty and
G
contains a single connected component ε.G (the graph itself), thus F0computes
F. On the other hand, if GGΣ×{1}×{1}then
Gis empty and
Gcontains
ε.G only, thus F0computes the identity. Hence this F0is a good candidate for
being a reversible extension of F. It remains now to check that F0is causal,
vertex-preserving and reversible.
[Causal] Shift-invariance, boundedness and continuity follow directly from the
shift-invariance, boundedness and continuity of both Fand the identity. [Reversible]
Replace Fby F1in the previous definition.
... Still it was shown that Reversible CA decompose as a finite-depth circuit of reversible, local gates [14,20,21]. The same holds true for Reversible CGD [11] in spite of the dynamicity of the neighbourhood relation. In the probabilistic setting, however, the implication fails [3,19]. ...
... All these K u will be local unitary operators and will commute with each other. To do so, we generalize the construction presented in [11]: the local update K u consists in applying the causal operator U , "putting aside" vertex u from the graph, and applying the inverse operator U † . ...
Preprint
Consider a graph having quantum systems lying at each node. Suppose that the whole thing evolves in discrete time steps, according to a global, unitary causal operator. By causal we mean that information can only propagate at a bounded speed, with respect to the distance given by the graph. Suppose, moreover, that the graph itself is subject to the evolution, and may be driven to be in a quantum superposition of graphs---in accordance to the superposition principle. We show that these unitary causal operators must decompose as a finite-depth circuit of local unitary gates. This unifies a result on Quantum Cellular Automata with another on Reversible Causal Graph Dynamics. Along the way we formalize a notion of causality which is valid in the context of quantum superpositions of time-varying graphs, and has a number of good properties. Keywords: Quantum Lattice Gas Automata, Block-representation, Curtis-Hedlund-Lyndon, No-signalling, Localizability, Quantum Gravity, Quantum Graphity, Causal Dynamical Triangulations, Spin Networks, Dynamical networks, Graph Rewriting.
... Still it was shown that Reversible CA decompose as a finite-depth circuit of reversible, local gates [14,18,19]. The same holds true for Reversible CGD [11] in spite of the dynamicity of the neighbourhood relation. In the probabilistic setting, however, the implication fails [3,17]. ...
... All these K u will be local unitary operators and will commute with each other. To do so, we generalize the construction presented in [11]: the local update K u consists in applying the causal operator U , "putting aside" vertex u from the graph, and applying the inverse operator U † . ...
Article
Consider a graph having quantum systems lying at each node. Suppose that the whole thing evolves in discrete time steps, according to a global, unitary causal operator. By causal we mean that information can only propagate at a bounded speed, with respect to the distance given by the graph. Suppose, moreover, that the graph itself is subject to the evolution, and may be driven to be in a quantum superposition of graphs---in accordance to the superposition principle. We show that these unitary causal operators must decompose as a finite-depth circuit of local unitary gates. This unifies a result on Quantum Cellular Automata with another on Reversible Causal Graph Dynamics. Along the way we formalize a notion of causality which is valid in the context of quantum superpositions of time-varying graphs, and has a number of good properties. Keywords: Quantum Lattice Gas Automata, Block-representation, Curtis-Hedlund-Lyndon, No-signalling, Localizability, Quantum Gravity, Quantum Graphity, Causal Dynamical Triangulations, Spin Networks, Dynamical networks, Graph Rewriting.
... This journal paper is based upon two conference proceedings (Arrighi et al. 2016(Arrighi et al. , 2015. It organized as follows. ...
Article
Full-text available
Causal Graph Dynamics extend Cellular Automata to arbitrary time-varying graphs of bounded degree. The whole graph evolves in discrete time steps, and this global evolution is required to have a number of symmetries: shift-invariance (it acts everywhere the same) and causality (information has a bounded speed of propagation). We add a further physics-like symmetry, namely reversibility. In particular, we extend two fundamental results on reversible cellular automata, by proving that the inverse of a causal graph dynamics is a causal graph dynamics, and that these reversible causal graph dynamics can be represented as finite-depth circuits of local reversible gates. We also show that reversible causal graph dynamics preserve the size of all but a finite number of graphs.
... In [7,6] one of the authors studied CGD in the reversible regime. Specific examples of these were described in [17,22]. ...
Preprint
Full-text available
Consider a network that evolves reversibly, according to nearest neighbours interactions. Can its dynamics create/destroy nodes?. On the one hand, since the nodes are the principal carriers of information, it seems that they cannot be destroyed without jeopardising bijectivity. On the other hand, there are plenty of global functions from graphs to graphs that are non-vertex-preserving and bijective. The question has been answered negatively---in three different ways. Yet, in this paper we do obtain reversible local node creation/destruction---in three relaxed settings, whose equivalence we prove for robustness. We motivate our work both by theoretical computer science considerations (reversible computing, cellular automata extensions) and theoretical physics concerns (basic formalisms for discrete quantum gravity).
Chapter
Consider a network that evolves reversibly, according to nearest neighbours interactions. Can its dynamics create/destroy nodes? On the one hand, since the nodes are the principal carriers of information, it seems that they cannot be destroyed without jeopardising bijectivity. On the other hand, there are plenty of global functions from graphs to graphs that are non-vertex-preserving and bijective. The question has been answered negatively—in three different ways. Yet, in this paper we do obtain reversible local node creation/destruction—in three relaxed settings, whose equivalence we prove for robustness. We motivate our work both by theoretical computer science considerations (reversible computing, cellular automata extensions) and theoretical physics concerns (basic formalisms towards discrete quantum gravity).
Conference Paper
Causal Graph Dynamics extend Cellular Automata to arbitrary, bounded-degree, time-varying graphs. The whole graph evolves in discrete time steps, and this global evolution is required to have a number of physics-like symmetries: shift-invariance (it acts everywhere the same) and causality (information has a bounded speed of propagation). We study a further physics-like symmetry, namely reversibility. We extend a fundamental result on reversible cellular automata by proving that the inverse of a causal graph dynamics is a causal graph dynamics. We also address the question of the evolution of the structure of the graphs under reversible causal graph dynamics, showing that any reversible causal graph dynamics preserves the size of all but a finite number of graphs.
Article
Cellular Automata constitute one of the most established model of discrete physical transformations that accounts for euclidean space. They implement three fundamental symmetries of physics: causality, homogeneity and finite density of information. Even though their origins lies in physics, they are widely used to model spatially distributed computation (self-replicating machines, synchronization problems,...), as well as a great variety of multi-agents phenomena (traffic jams, demographics,...). While being one of the most studied model of distributed computation, their rigidity forbids any trivial extension toward time-varying topology, which is a fundamental requirement when it comes to modelling phenomena in biology, sociology or physics: for instance when looking for a discrete formulation of general relativity. Causal graph dynamics generalize cellular automata to arbitrary, bounded degree, time-varying graphs. In this work, we generalize the fundamental structure results of cellular automata for this type of transformations. We endow our graphs with a compact metric space structure, and follow two approaches. An axiomatic approach based on the notions of continuity and shift-invariance, and a constructive approach, where a local rule is applied synchronously on every vertex of the graph. Compactness allows us to show the equivalence of these two definitions, extending the famous result of Curtis-Hedlund-Lyndon’s theorem. Another physics-inspired symmetry is then added to the model, namely reversibility.
Article
Full-text available
Parallel and distributed derivations are introduced and studied in the single-pushout approach, which models rewriting by pushout constructions in appropriate categories of partial morphisms. We present a categorical framework for this approach in an axiomatic way. Models of this categorical framework are among others: graphs, hypergraphs, relational structures, and algebraic specifications with suitable partial morphisms. Several new results concerning parallelism and distributed parallelism are presented which are even new in the example categories.
Article
Full-text available
The theorem of Ax says that any regular selfmapping of a complex algebraic variety is either surjective or non-injective; this property is called surjunctivity and investigated in the present paper in the category of proregular mappings of proalgebraic spaces. We show that such maps are surjunctive if they commute with sufficiently large automorphism groups. Of particular interest is the case of proalgebraic varieties over infinite graphs. The paper intends to bring out relations between model theory, algebraic geometry, and symbolic dynamics.
Article
Full-text available
The single-pushout approach to graph transformation interprets a double-pushout transformation rule of the classical algebraic approach which consists of two total graph morphisms as a single partial morphism from the left- to the right-hand side. The notion of a double-pushout diagram for the transformation process can then be substituted by a single-pushout diagram in an appropriate category of partial morphisms.It can be shown that this kind of transformation generalizes the double-pushout framework. Hence, the classical approach can be seen as a special (and very important) case of the new concept. It can be reobtained from the single-pushout approach by imposing an application condition on the redices which formulates the gluing conditions in the new setting. On the other hand, single-pushout transformations are always possible even if the gluing conditions for the redex are violated.The simpler structure of a direct transformation (one pushout diagram instead of two) simplifies many proofs. Hence, the whole theory for double-pushout transformations including sequential composition, parallel composition, and amalgamation can be reformulated and generalized in the new framework.Some constructions provide new effects and properties which are discussed in detail.
Article
The simplectic approximation to Einstein's equations ("Regge calculus") is derived by considering the net to be actually a (singular) Riemannian manifold. Specific nets for open and closed spaces are introduced in terms of which one can formulate the general time-evolution problem, which thereby reduces to the repeated solution of finite sets of coupled nonlinear (algebraic) equations. The initial-value problem is also formulated in simplectic terms.
Article
Cayley graphs have a number of useful features: the ability to graphically represent finitely generated group elements and their equality; to name all vertices relative to a point; the fact that they have a well-defined notion of translation, and that they can be endowed with a compact metric. We propose a notion of graph associated to a language, which conserves or generalizes these features. Whereas Cayley graphs are regular; associated graphs are arbitrary, although of a bounded degree. Moreover, it is well-known that cellular automata can be characterized as the set of translation-invariant continuous functions for a distance on the set of configurations that makes it a compact metric space; this point of view makes it easy to extend their definition from grids to Cayley graphs. Similarly, we extend their definition to these arbitrary, bounded degree, time-varying graphs. KEYWORDS: Causal Graph Dynamics, Curtis-Hedlund-Lynden, Dynamical networks, Boolean networks, Generative networks automata, Graph Automata, Graph rewriting automata, Parallel graph transformations, Amalgamated graph transformations, Time-varying graphs, Regge calculus, Local, No-signalling.
Article
We demonstrate the structural invertibility of all reversible one- and two-dimensional cellular automata. More precisely, we prove that every reversible two-dimensional cellular automaton can be expressed as a combination of four block permutations, and some shift-like mappings. Block permutations are very simple functions that uniformly divide configurations into rectangular regions of equal size and apply a fixed permutation on all regions.
Chapter
We introduce a framework called graph-rewriting automata to model evolution processes of networks. It is a natural extension of cellular automata in the sense that a fixed lattice space of cellular automata is extended to a dynamic graph structure by introducing local graph-rewriting rules. We consider three different constructions of rule sets to show that various network evolution is possible: hand-coding, evolutionary generation, and exhaustive search. Graph-rewriting automata provide a new tool to describe various complex systems and to approach many scientific problems.
Article
In the present paper we generalize the well-known Parallelism Theorem for graph derivations to the Amalgamation Theorem. In this theorem the assumption of “parallel independence” is dropped. For each pair of productions together with a relational production (allowing productions to be associated with each other) we construct a single “amalgamated” production. The Amalgamation Theorem states that graph derivations which respect the given associations can be amalgamated to a single derivation via the “amalgamated” production. The amalgamation of graph transformations can be considered as a synchronization mechanism. The amalgamation concept is applied to synchronization of graph manipulations in a simplified railway control system as well as to a graph grammar formalism for distributed systems (GDS).
Article
A variety of models of self-reproduction process have been proposed since von Neumann initiated this field with his self-reproducing automata. Almost all of them are described within the framework of two-dimensional cellular automata. They are heavily dependent on or limited by the peculiar properties of the two-dimensional lattice spaces. But such properties are irrelevant to the essential nature of self-replication. In this paper, we introduce a new framework called “graph automata” to obtain a natural description of complicated spatio-temporal developmental processes such as self-reproduction. The most advantageous point of this methodology is that it is not restricted to particular lattice space. As an illustrative example, a self-reproduction of Turing machine, which requires very long description by conventional cellular automata, is shown in a simple and straightforward formulation. Graph automata provide a new tool to approach important scientific problems such as evolution of morphology, and also to give the basis of self-reproducing and self-repairing artifacts.