Next Article in Journal
A Mathematical Method for Determining the Parameters of Functional Dependencies Using Multiscale Probability Distribution Functions
Previous Article in Journal
Mathematical Modelling of the Impact of Non-Pharmacological Strategies to Control the COVID-19 Epidemic in Portugal
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Orbit Entropy and Symmetry Index Revisited

by
Maryam Jalali-Rad
1,
Modjtaba Ghorbani
1,*,
Matthias Dehmer
2,3,4 and
Frank Emmert-Streib
5,6
1
Department of Mathematics, Faculty of Science, Shahid Rajaee Teacher Training University, Tehran 16785-136, Iran
2
Department of Computer Science, Swiss Distance University of Applied Sciences, 3900 Brig, Switzerland
3
Department of Biomedical Computer Science and Mechatronics, UMIT, 6060 Hall in Tyrol, Austria
4
College of Artificial Intelligence, Nankai University, Tianjin 300350, China
5
Predictive Society and Data Analytics Lab, Tampere University, Tampere, Korkeakoulunkatu 10, 33720 Tampere, Finland
6
Institute of Biosciences and Medical Technology, Tampere University, Tampere, Korkeakoulunkatu 10, 33720 Tampere, Finland
*
Author to whom correspondence should be addressed.
Mathematics 2021, 9(10), 1086; https://doi.org/10.3390/math9101086
Submission received: 23 March 2021 / Revised: 24 April 2021 / Accepted: 29 April 2021 / Published: 12 May 2021

Abstract

:
The size of the orbits or similar vertices of a network provides important information regarding each individual component of the network. In this paper, we investigate the entropy or information content and the symmetry index for several classes of graphs and compare the values of this measure with that of the symmetry index of certain graphs.

1. Introduction

Graph entropy measures were first introduced in the study of biological and chemical systems, with Rashevsky [1] and Mowshowitz [2,3,4,5] making the main contributions. In particular, Mowshowitz [5] interpreted the topological information content of a graph, such as its entropy measure. Since then, various graph entropy measures have been defined to investigate the structural properties of graphs [6,7,8] as well as [9,10,11,12,13].
A small-world graph [14,15] is a special type of graph in which the neighbors of any given vertex are likely to be neighbors of each other, but the probability that a vertex is the neighbor of another one is low and most vertices can be reached from each other by a few steps.
Adaptive networks are suitable to model the complex treatment represented by various real-world systems as well as to carry out decentralized information processing tasks such as drifting conditions and learning from online streaming data, see [16]. On the other hand, signal processing on graphs extends concepts and techniques from traditional signal processing to data indexed by generic graphs, see [17]. For example, neural networks and graph signal processing have emerged as important actors in data-science applications dealing with complex datasets, see [18].
This paper has two objectives. In Section 1, we investigate the automorphism group of some classes of graphs and verify their entropies and symmetry indices. In this way, some practical graph automorphism group decompositions are created that constitute the whole structure of graph automorphism groups.
In Section 2, we state concepts we use tp perform our analysis. We prove that there are several classes of graphs whose symmetry index is greater or equal than the orbit-entropy measure, while many other classes have a greater orbit entropy.

2. Entropy Measure and Symmetry Index of Graphs

Let G = ( V ( G ) , E ( G ) ) be a connected graph. An automorphism is a permutation α on the set of vertices of G with the property that both α and α 1 preserve the vertex adjacency. In other words, for two vertices u , v V ( G ) a permutation α on V ( G ) is an automorphism, when u v E ( G ) if and only if α ( u ) α ( v ) E ( G ) , where α ( u ) is the image of vertex u. The set of all automorphisms under the composition of maps forms a group denoted by Aut ( G ) .
For example consider the cycle graph C 3 in Figure 1. The line of symmetry colored by red is denoted by permutation ( 2 , 3 ) . Hence, the permutations correspond to blue and green lines are respectively denoted by ( 1 , 2 ) and ( 1 , 3 ) . A clockwise rotation equal 120 around the middle point of C 3 is denoted by ( 1 , 2 , 3 ) and equal 240 by ( 1 , 3 , 2 ) . All of these permutations preserve the vertex adjacency and thus are automorphisms. Note that a 360 rotation preserves the figure unchanged and we denote this permutation by ( ) . Hence, the automorphism group of C 3 has 6 elements which is A u t ( C 3 ) = { ( ) , ( 1 , 2 ) , ( 1 , 3 ) , ( 2 , 3 ) , ( 1 , 2 , 3 ) , ( 1 , 3 , 2 ) } .
For any vertex u V ( G ) an orbit of G containing u is defined as u G = O ( u ) = { α ( u ) : α A u t ( G ) } . We say G is vertex-transitive if it has only one orbit. Equivalently, a graph is vertex-transitive if for two vertices u , v V ( G ) there is an automorphism σ A u t ( G ) such that σ ( u ) = v .
The orbits of the automorphism group of a graph form a partition of the vertices of the graph. This decomposition introduces the symmetry structure of the graph, and the orbit entropy measure obtained from the automorphism group provides an index of the complexity of the graph relative to the symmetry structure.
Mowshowitz [2] defined the topological information content, which is a classical graph entropy measure, as
I a ( G ) = i = 1 k | O i | | V | log ( | O i | | V | ) ,
where O i ( 1 i k ) are orbits of G under the action of automorphism group on the set of vertices. The collection of k orbits { O 1 , , O k } defines a finite probability scheme in an obvious way. This measure is addressed to the problem of measuring the relative complexity of graphs. The idea of measuring the information content of a graph was first presented in [1]. Mowshowitz and Dehmer [19] defined the symmetry index S ( G ) as
S ( G ) = ( log n I a ( G ) ) + log | Aut ( G ) | = 1 n ( i = 1 k | O i | log | O i | ) + log | Aut ( G ) | .

2.1. Relationship between Symmetry Index and Orbit Entropy

Consider a permutation σ on the set X = { x 1 , , x n } . Then the set of all elements that σ moves is called the support of σ . Two permutations σ and γ are disjoint if their supports have no intersection. Consider S to be a set of generators of Aut ( G ) , e S and S = S 1 S m to be the partition of S, where S i cannot be decomposed into smaller support-disjoint subsets. Therefore, we have the following.
Theorem 1.
Ref. [20] If A i = S i , then Aut ( G ) A 1 × A 2 × × A m .
Consider A = Aut ( G ) . For a vertex v in V ( G ) , A v denotes the stabilizer subgroup of G containing all automorphisms that fix the vertex v. Similarly, for a vertex u of G, its orbit is a set containing all α ( u ) , where α is an automorphism of A u t ( G ) .
Theorem 2.
Ref. [21] (Orbit-stabilizer Theorem) Let A be a permutation group acting on a set Ω and u be an arbitrary point in the set Ω. Then | A | = | A u | | u A | .
Definition 1.
Harary [22] defined the corona product G 1 G 2 of two graphs G 1 and G 2 as a new graph G obtained by taking one copy of G 1 (which has p 1 vertices) together with p 1 copies of G 2 and then joining the ith vertex of G 1 to all vertices in the ith copy of G 2 ; see Figure 2.
Suppose A and B are two finite groups in which B acts on the set X. The wreath product of A and B (denoted by A B ) is a group with the underline set
A B = { ( f ; h ) | f : X A i s a f u n c t i o n , h B } .
The group operation can be defined as ( f 1 ; h 1 ) ( f 2 ; h 2 ) = ( g ; h 1 h 2 ) , where for each element i X , we obtain
g ( i ) = f 1 ( i ) f 2 ( i h 1 ) .
Wreath product is one of the most significant combinatorial buildings in the field of permutation group theory. The next theorem shows that the automorphism group of a big graph can be constructed from wreath product of automorphism groups of its subgraphs.
Theorem 3.
Ref. [22] The automorphism group of the corona product G 1 G 2 of two graphs is isomorphic to the wreath product Aut ( G 1 ) Aut ( G 2 ) if and only if either G 1 or G 2 has no isolated vertices.
Let G and H be two disjoint graphs, where V ( G ) = { u 1 , , u n } , and let u V ( G ) and v V ( H ) . The splice of two graphs G and H by vertices u and v, denoted by G H ( u , v ) , is a new graph constructed by identifying two vertices u and v in the union of G and H [23]. Similarly, let K be a graph constructed by G and n copies of graph H and then splicing vertex u i of G by vertex v of the ith copy of H. The following result is contained in [24,25,26,27,28].
Theorem 4. (Balasubramanian)
Let K be as defined above. Then Aut ( K ) Aut ( G ) Aut ( H ) .
Here, the orbit-entropy measure and the symmetry index of some classes of well-known graphs are determined, followed by a comparison of these measures for all the graphs. The results show that whether a measure is greater or smaller cannot be predicted in advance, and it depends on the structure of the graph or equivalently the structure of its automorphism group.
Example 1.
It is demonstrated that the orbit entropy of the path graph P n is greater than its symmetry index, whereas for the wheel graph W n , the star graph S n , and the complete bipartite graph K m , n , the orbit entropy is less than the symmetry index. To do this, consider the following cases:
(a) 
If n is odd, then Aut ( P n ) Z 2 and P n has n 1 2 orbits of size two and a singleton orbit. Hence, S ( P n ) = 2 n 1 n and I a ( P n ) = log n n 1 n . This means that if n 7 , then I a ( P n ) > S ( P n ) .
(b) 
If n is even, then Aut ( P n ) Z 2 and G has n / 2 orbits of order 2. Hence, S ( G ) = 2 and I a ( G ) = log n 1 .
(c) 
If G = S n , then Aut ( G ) S n 1 and G has a singleton orbit together with an orbit of size n 1 . Thus,
S ( G ) = n 1 n log ( n 1 ) + log ( ( n 1 ) ! )
and
I a ( G ) = log n n 1 n log ( n 1 ) .
(d) 
For the wheel graph W n , it is well known that Aut ( W n ) D 2 ( n 1 ) , and consequently it has a singleton orbit and an orbit of size n 1 . Hence,
S ( W n ) = 2 n 1 n log ( n 1 ) + 1 and I a ( W n ) = I a ( S n ) .
(e) 
If G = K m , n , then Aut ( G ) S n × S m , and consequently G has two orbits of sizes m and n. Thus,
S ( G ) = m n + m log m + n n + m log n + log ( m ! n ! )
and
I a ( G ) = log ( m + n ) m n + m log ( m n + m ) n n + m log ( n n + m ) .
This completes the proof.
Theorem 5.
Let G be a graph with an automorphism group containing the identity element alone. Then I a ( G ) > S ( G ) .
Proof. 
Assume that G is a graph on n vertices and Aut ( G ) id . Hence, G has n orbits, giving
S ( G ) = 1 n i = 1 n 1 log 1 + log 1 = 0
and I a ( G ) = log n . □
If the automorphism group of G acts transitively on V ( G ) , then it is concluded that G is vertex-transitive. Equivalently, a vertex-transitive graph has only one orbit. Similarly, an edge-transitive graph can be defined.
Theorem 6.
If G is vertex-transitive, then I a ( G ) < S ( G ) .
Proof. 
Because G has only one orbit, S ( G ) = log ( n | Aut ( G ) | ) and I a ( G ) = n n log 1 = 0 . Therefore, I a ( G ) < S ( G ) . □
The Cayley graph G = C a y ( A , S ) is a graph constructed from a group A and a subset S A , where e S and S 1 = S . The vertex set of graph G comprises the elements of A, and two vertices a and b are adjacent if and only if b 1 a S .
Corollary 1.
If G is a Cayley graph or G C n , K n , then the symmetry index of G is greater than its orbit-entropy measure.
Proof. 
It is well known that each Cayley graph is vertex-transitive [29], and the assertion follows. On the other hand, it is well known that Aut ( C n ) D 2 n , where D 2 n denotes the dihedral group of order 2 n with the following presentation:
D 2 n = x , y : x n = y 2 = 1 , y 1 x y = x 1 .
Also, Aut ( K n ) S n , which implies that both C n and K n are vertex-transitive, and we are done. □
Two vertices u and v of a graph G are said to be similar if there is an automorphism α Aut ( G ) such that α ( u ) = v . Herein, all similar vertices have the same color.
Corollary 2.
For the MacPherson graph G = M ( s , t ) as shown in Figure 3, Aut ( M ( s , t ) ) S s S s 1 S s 1 S s 1 and S ( M ( s , t ) ) > I a ( M ( s , t ) ) .
Proof. 
It is obvious that G is constructed by K s and n copies of graph K s 1 and then by splicing each vertex v i of K s by vertex u of the ith copy of K s 1 . This means that Aut ( M ( s , t ) ) S s S s 1 S s 1 S s 1 . In [30], it is proved that G is vertex-transitive. Considering this fact and Theorem 6, we may conclude the assertion. □
Theorem 7.
If G is a regular edge-transitive graph, then I a ( G ) < S ( G ) .
Proof. 
Note first that if G is vertex-transitive, then I a ( G ) < S ( G ) from Theorem 6. Thus, we can assume that G is not vertex-transitive. Therefore, it is bipartite, and because G is regular, two independent sets are exactly the orbits of Aut ( G ) on V ( G ) . This means that G K m , m , and by Example 1, the required result is obtained. □
Theorem 8.
Let G be a graph of order n 3 , with n 2 singleton orbits and an orbit of order two. Then I a ( G ) S ( G ) , and for n = 3 , we have S ( G ) > I a ( G ) . In addition, if G has n 3 singleton orbits and an orbit of order three, then I a ( G ) S ( G ) ( n 8 ) .
Proof. 
Suppose that n = 3 and G has an orbit of size two. Then G P 3 and thus S ( G ) > I a ( G ) . If n > 3 , then Aut ( G ) Z 2 and so S ( G ) = n + 2 n , and I a ( G ) = log n 2 n . If n = 4 , then I a ( G ) = S ( G ) = 3 2 . If n 5 , then S ( G ) = n + 2 n 7 5 and I a ( G ) = log n 2 n 8 5 . This gives I a ( G ) > S ( G ) , and the assertion follows. If G has n 3 singleton orbits and an orbit of order three, then Aut ( G ) S 3 or Aut ( G ) Z 3 . If Aut ( G ) S 3 , then S ( G ) = ( n + 3 n ) log ( 3 ) and I a ( G ) = log n 3 n log 3 . One can see that for n 8 , S ( G ) log 9 and thus I a ( G ) > S ( G ) . If Aut ( G ) S 3 , then S ( G ) = 1 + n + 3 n log 3 and I a ( G ) = log n 3 n log 3 . Thus we have I a ( G ) S ( G ) for n 11 . □
Theorem 9.
Consider G to be a graph of order n, with n k singleton orbits and an orbit of order k. Consider the size of the automorphism group to be less than or equal to n k 2 k n . Then S ( G ) I a ( G ) .
Proof. 
The following arise by definition:
S ( G ) = log ( | Aut ( G ) | ) + k n = log ( | ( Aut ( G ) | k k n ) , I a ( G ) = log n + k n log k = log ( n k 2 k n ) .
Therefore, if | Aut ( G ) | n k 2 k n , then S ( G ) I a ( G ) . □
Theorem 10.
Assume that G is a graph on n vertices and that the orbits of A = Aut ( G ) are of equal size. Thus for n s , we have S ( G ) I a ( G ) , where v is an arbitrary vertex in V ( G ) and s = | A | 3 | A v | 2 .
Proof. 
Suppose that G has k orbits. Then S ( G ) = log ( m | A | ) and I a ( G ) = log ( n m ) = log k , where m is the orbit size. Thus, S ( G ) I a ( G ) if and only if m | Aut ( G ) | k if and only if n | A | 3 | A v | 2 = s . For a given automorphism σ of graph G, the fix point of σ is defined by f i x ( σ ) = { v V ( G ) : σ ( v ) = v } . □
Theorem 11.
Let G be a graph on n vertices and A = Aut ( G ) . If σ Aut ( G ) | f i x ( σ ) | < n a log ( | A | ) log n , then S ( G ) > I a ( G ) .
Proof. 
Suppose that O 1 , , O r are all orbits of A. Because | O i | 1 , we obtain | O i | n log | O i | n 1 n log 1 n , thus i = 1 r | O i | n log | O i | n r n log n . Hence, I a ( G ) r n log ( n ) . From the Burnside Lemma [21], the number of orbits is r = 1 | A | σ A | f i x ( σ ) | . This yields I a ( G ) 1 n | A | σ A | fix ( σ ) | log ( n ) . Now if σ A | f i x ( σ ) | < n a log ( | A | ) log n , then log ( | A | ) > log ( n ) n | A | σ A | f i x ( σ ) | . Thus, log ( | A | ) > I a ( G ) and so
1 n σ A | O i | log ( | O i | ) + log ( | A | ) > I a ( G ) + 1 n σ A | O i | log | O i | .
This leads to S ( G ) > I a ( G ) . □
Theorem 12.
Suppose that G 1 K n + { u } and G 2 K n + { e } ; see Figure 4. Then I a ( G 1 ) < S ( G 1 ) and I a ( G 2 ) < S ( G 2 ) .
Proof. 
Clearly, G 1 has n vertices of degree n 1 and a vertex of degree two. This leads to Aut ( G 1 ) S n 2 × Z 2 . Hence G 1 has three orbits of sizes one, two, and n 2 , respectively, which means that
S ( G 1 ) = n + 3 n + 1 + n 2 n + 1 log ( n 2 ) + log ( ( n 2 ) ! ) , I a ( G 1 ) = log ( n + 1 ) n 2 n + 1 log ( n 2 ) 2 n + 1 .
Meanwhile, G 2 has n 1 vertices of degree n 1 , a vertex of degree n, and a vertex of degree one. This together with the fact that Aut ( G 2 ) S n 1 leads us to assume that G 2 has three orbits of sizes one, one, and n 1 , respectively, and therefore
S ( G 2 ) = n 1 n + 1 log ( n 1 ) + log ( ( n 1 ) ! ) , I a ( G 2 ) = log ( n + 1 ) n 1 n + 1 log ( n 1 ) .
This completes the proof. □
A caterpillar tree is a tree in which all the vertices are within distance one of a central path. In other words, the caterpillar tree C ( n 1 , n 2 , , n r ) consists of a path with r vertices in which n i pendent edges are attached to the ith vertex of P r ; see Figure 5.
Theorem 13.
Let G = C ( n 1 , n 2 , , n r ) be a caterpillar tree, where n 1 n r . Then I a ( G ) < S ( G ) .
Proof. 
The graph G has i = 1 r n i + r = n vertices, and from Theorem 1 we have that Aut ( G ) S n 1 × S n 2 × × S n r . This leads to the conclusion that G has 2 r orbits, and thus
S ( G ) = 1 n i = 1 r n i log ( n i ) + log ( n 1 ! n r ! )
and
I a ( G ) = log n i = 1 r n i n log n i .
Hence, I a ( G ) < S ( G ) . □
Theorem 14.
If T is a tree with two orbits, then I a ( G ) < S ( G ) .
Proof. 
Let T be a tree with two orbits. It is well known that T S n or T B n , n [22]; see Figure 6. If T B n , n , then Aut ( G ) S n × S n × Z 2 , and thus S ( G ) = 2 log ( n ! ) + n n + 1 log ( n ) + 2 . Meanwhile, I a ( G ) = log ( n + 1 ) n n + 1 log ( n ) . Hence, I a ( B n , n ) < S ( B n , n ) . If T S n , then from Example 1 we obtain I a ( S n ) < S ( S n ) . □
Theorem 15.
If T is a tree with three orbits, then I a ( G ) < S ( G ) .
Proof. 
It is well known that each tree has either a central vertex or a central edge [31]. First, suppose that T has a central vertex. The central vertex gives rise to a singleton orbit, and the leaf vertices are in the second orbit. If the leaf vertices give rise to at least two orbits, then T has at least four orbits, which is a contradiction. Hence, the leaf vertices necessarily lie in the same orbit. The other vertices are in the same orbit, therefore they all have the same degree. On the other hand, the leaf vertices are adjacent to the central vertex because there are only three orbits. This leads us to investigate T T n , r ; see Figure 7. Clearly, T has 1 + n + n r vertices, and Theorem 4 gives Aut ( T ) S n S r , where the sizes of the orbits of T are one, n, and n r , respectively. This means that
S ( T ) = n ( 1 + r ) n ( 1 + r ) + 1 log n + n r n ( 1 + r ) + 1 log r + log ( n ! r ! n ) , I a ( T ) = log ( n ( 1 + r ) + 1 ) n ( 1 + r ) n ( 1 + r ) + 1 log n n r n ( 1 + r ) + 1 log r .
Assume that T has a central edge. By a similar argument, it can be proved that T is isomorphic with graph D T n , n , as shown in Figure 8. Theorem 4 gives Aut ( T ) ( S 2 S n ) S r , and thus T has three orbits of sizes two, 2 n , and 2 n r , respectively. Therefore, we have that
S ( T ) = 2 log ( n ( 1 + r ) + 1 ) 2 n ( 1 + r ) n ( 1 + r ) + 1 log n 2 n r n ( 1 + r ) + 1 log r + 2 , I a ( T ) = log ( n ( 1 + r ) + 1 ) n ( 1 + r ) n ( 1 + r ) + 1 log n n r n ( 1 + r ) + 1 log r .
This completes the proof. □

2.2. Orbit Entropy and Symmetry Index of Dendrimers

A dendrimer is a molecular graph associated with a dendrimer molecule. In this section, we determine the entropy or information content of some dendrimers. In the case of organic molecules, the lower the information content (or the greater the symmetry), the fewer the possibilities for different interactions with other molecules. If all the atoms are in the same equivalence classes, then it makes no difference which one interacts with an atom of another molecule.
Theorem 16.
For the dendrimer G n shown in Figure 9, we obtain S ( G n ) > I a ( G n ) .
Proof. 
The fact that
| V ( G n ) | = 6 + 4 i = 2 n + 1 2 i = 6 + 4 ( 2 n + 2 4 ) = 2 n + 4 10
can be verified from Figure 9 and Figure 10. It is not difficult to prove that Aut ( G n ) Aut ( G n 1 ) Z 2 . Using induction on n yields Aut ( G n ) Z 2 : ( Z 2 × Z 2 × × Z 2 ( 2 n + 1 2 ) times ) , where “:” denotes the semi-direct product [22]. It is clear that G n has three orbits of size two and four orbits of size 2 m , 2 m n + 1 . Hence, the symmetry index and the orbit-entropy measure of dendrimer G n are given by
S ( G n ) = log ( 2 n + 2 4 ) + 6 + m = 2 n + 1 m 2 m + 2 2 n + 4 10 , I a ( G n ) = ( 6 2 n + 4 10 log ( 2 2 n + 4 10 ) + m = 2 n + 1 2 m + 2 2 n + 4 10 log ( 2 m 2 n + 4 10 ) ) = log ( 2 n + 4 10 ) 6 + m = 2 n + 1 m 2 m + 2 2 n + 4 10 .
Here, we study the orbit-entropy measure and the symmetry index of another class of dendrimers, namely H n , shown in Figure 11. For a given vertex v V ( G ) , suppose that N G ( v ) = { u V ( G ) ; u v V ( G ) } and X is a subset of vertices of graph G. By X , we mean the induced subgraph of G with vertex set X, and two vertices in X are adjacent if and only if they are adjacent in G. The central vertex t has degree three and N G ( w ) = { x 0 , y 0 , z 0 } . There are three branches rooted at vertices x 0 , y 0 , and z 0 . For two vertices u , v V ( G ) , the distance between them is the length of the shortest path connecting them, denoted by d ( u , v ) . By the ith level of H n , we mean the set of vertices at distance i from the central vertex w.
Theorem 17.
For the dendrimer graph H n , we have S ( H n ) > I a ( H n ) .
Proof. 
It is obvious that | V ( H n ) | = 4 + 3 i = 1 n 1 2 i = 3 ( 2 n ) 2 . Similarly, Aut ( H 1 ) S 3 , and if n 2 , then we have Aut ( H n ) Aut ( H n 1 ) Z 2 from the structure of H n . Now, using induction on n shows that Aut ( H n ) S 3 : ( Z 2 × Z 2 × × Z 2 3 ( 2 n 1 1 ) times ) .
Hence, H n has n + 1 orbits as follows. O 1 is a singleton set containing the central vertex w. For i 2 , the ith orbit contains all vertices at distance i from the central vertex. Hence, the vertices of the kth level of this graph are the vertices of the form a = x w , a = y w , or a = z w of length k, where w { 0 , 1 } k 1 . Therefore, we have that
S ( H n ) = log ( 3 2 ( 2 n 2 ) ) + 1 3 ( 2 n ) 2 i = 0 n 1 3 ( 2 i ) log 3 ( 2 i ) ,
I a ( H n ) = ( i = 0 n 1 3 ( 2 i ) 3 ( 2 n ) 2 log ( 3 ( 2 i ) 3 ( 2 n ) 2 ) ) + 1 3 ( 2 n ) 2 log ( 1 3 ( 2 n ) 2 ) ) = log ( 3 ( 2 n ) 2 ) 1 3 ( 2 n ) 2 i = 0 n 1 3 ( 2 i ) log ( 3 ( 2 i ) ) .
Consider the lattice graph L ( m , n ) that is the Cartesian product of two graphs P n and P m , see Figure 12. Thus, we have the following.
Theorem 18.
Let G = L ( m , n ) . If n = m , then we have S ( G ) < I a ( G ) for n 21 . If n m , then we have S ( G ) < I a ( G ) for ( m + 1 ) ( n + 1 ) 64 .
Proof. 
  • Consider m = n . Thus, G has ( n + 1 ) 2 vertices and A u t ( G ) D 8 , and the following two cases hold.
    (a)
    If n is odd, then G has n + 1 2 orbits of size four and n 2 1 8 orbits of size eight. Thus, S ( G ) = 6 n + 4 n + 1 and I a ( G ) = 2 log ( n + 1 ) 3 n + 1 n + 1 . For n 21 , we obtain log ( n + 1 ) > 9 n + 7 2 ( n + 1 ) , thus S ( G ) < I a ( G ) .
    (b)
    If n is even, then G has n orbits of size four, n ( n 2 ) 8 orbits of size eight, and a singleton orbit. Hence, S ( G ) = 6 n 2 + 8 n + 3 ( n + 1 ) 2 and I a ( G ) = 2 log ( n + 1 ) 3 n 2 + 2 n ( n + 1 ) 2 . For n 20 , log ( n + 1 ) > 9 n 2 + 10 n + 3 2 ( n + 1 ) 2 , and thus S ( G ) < I a ( G ) .
  • Consider m n . Thus, A u t ( G ) = Z 2 × Z 2 , and the following four cases hold.
    (a)
    If m and n are odd, then G has ( m + 1 ) ( n + 1 ) 4 orbits of size four. Thus, S ( G ) = 4 and I a ( G ) = log ( ( m + 1 ) ( n + 1 ) ) 2 . Therefore, if ( m + 1 ) ( n + 1 ) 64 , then S ( G ) I a ( G ) .
    (b)
    If m is even and n is odd, then G has m ( n + 1 ) 4 orbits of size four and n + 1 2 orbits of size two. Thus, S ( G ) = 4 m + 3 m + 1 and I a ( G ) = log ( ( m + 1 ) ( n + 1 ) ) 2 m + 1 m + 1 . Therefore, if ( m + 1 ) ( n + 1 ) 64 , then S ( G ) < I a ( G ) .
    (c)
    If m is odd and n is even, then G has n ( m + 1 ) 4 orbits of size four and m + 1 2 orbits of size two. Thus, S ( G ) = 4 n + 3 n + 1 and I a ( G ) = log ( ( m + 1 ) ( n + 1 ) ) 2 n + 1 n + 1 . For ( m + 1 ) ( n + 1 ) 64 , we obtain S ( G ) < I a ( G ) .
    (d)
    If m and n are both even, then G has n m 4 orbits of size four, n + m 2 orbits of size two, and one singleton orbit. Thus, S ( G ) = 4 mn + 3 m + 3 n + 2 ( m + 1 ) ( n + 1 ) and
    I a ( G ) = log ( ( m + 1 ) ( n + 1 ) ) mn + m + n ( m + 1 ) ( n + 1 ) .
    Therefore, if ( m + 1 ) ( n + 1 ) 32 , then S ( G ) < I a ( G ) .

3. Summary and Conclusions

Quantitative measures of graph complexity, defined in terms of Shannon entropy, are often based on vertex partitions [2,7]. For instance, partitions of the vertices of a graph are related to symmetry structure if they are based on vertex orbit cardinalities. In this paper, we investigated the orbit entropy [2] and the symmetry index [19] for several classes of graphs. We compared the values of these measures based on inequalities. As a result, we found several classes of graphs whose symmetry index was greater or equal to the orbit-entropy measure, while many other classes had a greater orbit entropy.
We also obtained useful and novel measures based on automorphism group decompositions. These measures should be compared with other existing graph complexity measures, which is left as future work.

Author Contributions

M.J.-R., M.G., M.D. and F.E.-S. wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

Matthias Dehmer thanks the Austrian Science Funds for supporting this work (project P30031).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rashevsky, N. Life, information theory, and topology. Bull. Math. Biophys. 1955, 17, 229–235. [Google Scholar] [CrossRef]
  2. Mowshowitz, A. Entropy and the complexity of the graphs: I. An index of the relative complexity of a graph. Bull. Math. Biophys. 1968, 30, 175–204. [Google Scholar] [CrossRef]
  3. Mowshowitz, A. Entropy and the complexity of graphs II: The information content of digraphs and infinite graphs. Bull. Math. Biophys. 1968, 30, 225–240. [Google Scholar] [CrossRef]
  4. Mowshowitz, A. Entropy and the complexity of graphs III: Graphs with prescribed information content. Bull. Math. Biophys. 1968, 30, 387–414. [Google Scholar] [CrossRef]
  5. Mowshowitz, A. Entropy and the complexity of graphs IV: Entropy measures and graphical structure. Bull. Math. Biophys. 1968, 30, 533–546. [Google Scholar] [CrossRef]
  6. Ghorbani, M.; Dehmer, M.; Rajabi-Parsa, M.; Emmert-Streib, F.; Mowshowitz, A. Hosoya entropy of fullerene graph. Appl. Math. Comput. 2019, 352, 88–98. [Google Scholar]
  7. Dehmer, M.; Mowshowitz, A. A history of graph entropy measures. Inf. Sci. 2011, 181, 57–78. [Google Scholar] [CrossRef]
  8. Dehmer, M.; Shi, Y.; Emmert-Streib, F. Graph distance measures based on topological indices revisited. Appl. Math. Comput. 2015, 266, 623–633. [Google Scholar] [CrossRef]
  9. Mowshowitz, A.; Dehmer, M.; Emmert-Streib, F. A note on graphs with prescribed orbit structure. Entropy 2019, 21, 1118. [Google Scholar] [CrossRef] [Green Version]
  10. Ghorbani, M.; Dehmer, M.; Zangi, S. Graph operations based on using distance-based graph entropies. Appl. Math. Comput. 2018, 333, 547–555. [Google Scholar] [CrossRef]
  11. Ghorbani, M.; Dehmer, M. Properties of entropy-based topological measures of fullerenes. Mathematics 2020, 8, 740. [Google Scholar] [CrossRef]
  12. Chen, Y.; Zhao, Y.; Han, X. Characterization of symmetry of complex networks. Symmetry 2019, 11, 692. [Google Scholar] [CrossRef] [Green Version]
  13. Machado, J.A.T. An evolutionary perspective of virus propagation. Mathematics 2020, 8, 779. [Google Scholar] [CrossRef]
  14. Ma, F.; Yao, B. A family of small-world network models built by complete graph and iteration-function. Phys. Stat. Mech. Appl. 2018, 492, 2205–2219. [Google Scholar] [CrossRef]
  15. Chalupa, D.; Blum, C. Mining k-reachable sets in real-world networks using domination in shortcut graphs. J. Comput. Sci. 2017, 22, 1–14. [Google Scholar] [CrossRef] [Green Version]
  16. Djuric, P.M.; Richard, C. (Eds.) Cooperative and Graph Signal Processing; Academic Press: Cambridge, MA, USA, 2018. [Google Scholar]
  17. Sandryhaila, A.; Moura, J.M.F. Discrete signal processing on graphs: Frequency analysis. IEEE Trans. Signal Process. 2014, 62, 3042–3054. [Google Scholar] [CrossRef] [Green Version]
  18. Rey, S.; Tenorio, V.; Rozada, S.; Martino, L.; Marques, A.G. Deep Encoder-Decoder Neural Network Architectures for Graph Output Signals. In Proceedings of the IEEE Conference on Signals, Systems, and Computers (ASILOMAR), Pacific Grove, CA, USA, 3–6 November 2019. [Google Scholar]
  19. Mowshowitz, A.; Dehmer, M. A symmetry index for graphs. Symmetry Cult. Sci. 2010, 21, 321–327. [Google Scholar]
  20. MacArthur, B.D.; Sanches-Garcia, R.J.; Anderson, J.W. Symmetry in complex graphs. Discret. Appl. Math. 2008, 156, 3525–3531. [Google Scholar]
  21. Dixon, J.D.; Mortimer, B. Permutation Groups; Springer: New York, NY, USA, 1996. [Google Scholar]
  22. Harary, F. Graph Theory; Addison-Wesley Publishing Company: Boston, MA, USA, 1969. [Google Scholar]
  23. Došlić, T. Splices, links and their degree-weighted Wiener polynomials. Graph Theory Notes N. Y. 2005, 48, 47–55. [Google Scholar]
  24. Balasubramanian, K. Computer perception of NMR symmetrys. J. Magnet. Reson. 1995, 112, 182–190. [Google Scholar] [CrossRef]
  25. Balasubramanian, K. Generators of the character tables of generalized wreath product groups. Theor. Chim. Acta 1990, 78, 31–43. [Google Scholar] [CrossRef]
  26. Balasubramanian, K. Symmetry operators of generalized wreath products and their applications to chemical physics. Int. J. Quantum Chem. 1982, 22, 1013–1031. [Google Scholar] [CrossRef] [Green Version]
  27. Balasubramanian, K. The symmetry groups of nonrigid molecules as generalized wreath products and their representations. J. Chem. Phys. 1980, 72, 665–677. [Google Scholar] [CrossRef]
  28. Tan, H.; Liao, M.Z.; Balasubramanian, K. A flexible correlation group table (CGT) method for the relativistic configuration wave functions. J. Math. Chem. 2000, 28, 213–239. [Google Scholar] [CrossRef]
  29. Biggs, N. Algebraic Graph Theory; Cambridge University Press: London, UK, 1974. [Google Scholar]
  30. Cameron, P.J. Permutation Groups; London Mathematical Society Student Texts; Cambridge University Press: Cambridge, UK, 1999. [Google Scholar]
  31. Ghorbani, M.; Dehmer, M.; Mowshowitz, A.; Tao, J.; Emmert-Streib, F. The Hosoya entropy of graphs revisited. Symmetry 2019, 11, 1013. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The automorphism group A u t ( C 3 ) contains six permutations.
Figure 1. The automorphism group A u t ( C 3 ) contains six permutations.
Mathematics 09 01086 g001
Figure 2. G 1 , G 2 , G 1 G 2 , and G 2 G 1 .
Figure 2. G 1 , G 2 , G 1 G 2 , and G 2 G 1 .
Mathematics 09 01086 g002
Figure 3. MacPherson graph M ( 4 , 2 ) .
Figure 3. MacPherson graph M ( 4 , 2 ) .
Mathematics 09 01086 g003
Figure 4. Two graphs K n + u and K n + e for n = 7 .
Figure 4. Two graphs K n + u and K n + e for n = 7 .
Mathematics 09 01086 g004
Figure 5. Caterpillar graph C ( 2 , 1 , 0 , 3 , 1 , 4 ) .
Figure 5. Caterpillar graph C ( 2 , 1 , 0 , 3 , 1 , 4 ) .
Mathematics 09 01086 g005
Figure 6. Graph B m , n for m = 3 and n = 7 .
Figure 6. Graph B m , n for m = 3 and n = 7 .
Mathematics 09 01086 g006
Figure 7. Tree T n , r for n = 3 and r = 2 .
Figure 7. Tree T n , r for n = 3 and r = 2 .
Mathematics 09 01086 g007
Figure 8. The tree D T 3 , 5 in Theorem 15 for n = 3 and r = 5 . It has a central edge. Vertices in an orbit have the same color.
Figure 8. The tree D T 3 , 5 in Theorem 15 for n = 3 and r = 5 . It has a central edge. Vertices in an orbit have the same color.
Mathematics 09 01086 g008
Figure 9. Dendrimer G 1 in Theorem 16. Vertices in an orbit have the same color.
Figure 9. Dendrimer G 1 in Theorem 16. Vertices in an orbit have the same color.
Mathematics 09 01086 g009
Figure 10. Dendrimer G 2 .
Figure 10. Dendrimer G 2 .
Mathematics 09 01086 g010
Figure 11. Dendrimer H 4 .
Figure 11. Dendrimer H 4 .
Mathematics 09 01086 g011
Figure 12. Graph L ( 4 , 4 ) .
Figure 12. Graph L ( 4 , 4 ) .
Mathematics 09 01086 g012
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jalali-Rad, M.; Ghorbani, M.; Dehmer, M.; Emmert-Streib, F. Orbit Entropy and Symmetry Index Revisited. Mathematics 2021, 9, 1086. https://doi.org/10.3390/math9101086

AMA Style

Jalali-Rad M, Ghorbani M, Dehmer M, Emmert-Streib F. Orbit Entropy and Symmetry Index Revisited. Mathematics. 2021; 9(10):1086. https://doi.org/10.3390/math9101086

Chicago/Turabian Style

Jalali-Rad, Maryam, Modjtaba Ghorbani, Matthias Dehmer, and Frank Emmert-Streib. 2021. "Orbit Entropy and Symmetry Index Revisited" Mathematics 9, no. 10: 1086. https://doi.org/10.3390/math9101086

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop