1. Introduction
The fusion of algebraic concepts with graph-theoretic frameworks has established a vibrant interdisciplinary domain, wherein algebraic structures are analyzed using graph invariants and vice versa. Among such developments, a particularly fertile area involves associating graphs to commutative rings in order to gain insights into the ring’s internal behavior via its graphical representation. In these constructions, ring elements or substructures such as ideals are designated as vertices, and adjacency is defined by algebraic conditions that encode interaction, divisibility, or annihilation properties. This algebra–graph theory interface has matured into a significant research strand, contributing to both structural ring theory and combinatorics.
A seminal idea in this context is the concept of the zero-divisor graph, originally formulated by Beck [
1], who proposed using vertex adjacency to model zero-product interactions for investigating ring coloring problems. This idea was rigorously reformulated by Anderson and Livingston [
2], who standardized the notation and definition. For a commutative ring
R, the zero-divisor graph
is defined on the set
, where two distinct vertices,
x and
y, are connected if and only if
. Such graphs encode annihilation behavior in a simple combinatorial structure, making them a powerful vehicle for analyzing ring-theoretic properties.
Over recent decades, zero-divisor graphs have been extensively investigated. Lucas studied their diameters and addressed how the underlying ring structure influences distance parameters [
3]. Redmond [
4] examined finite examples of small cardinality, classifying the associated graphs and exploring extremal cases. Maimani et al. [
5] explored the topological genus of such graphs, thereby connecting ring theory to graph embeddings and surface topology. These works, among others, underscore how graph-theoretic descriptors like connectivity, genus, and chromatic number serve as indirect tools for probing algebraic frameworks.
The rise of spectral graph theory has added an analytic dimension to the study of graphs associated with rings. By analyzing the eigenvalues of matrices associated with such graphs—including adjacency, Laplacian, and distance Laplacian matrices—researchers have derived spectral signatures that correlate with the ring’s internal algebraic characteristics. In this spirit, normalized Laplacian spectra of weakly zero-divisor graphs [
6], and Randić and distance Laplacian spectra for zero-divisor graphs [
7,
8], have provided finer distinctions between ring classes and motivated deeper combinatorial analysis.
Beyond zero-divisor graphs, another rich construction is the total graph of a commutative ring. Introduced by Anderson and Badawi [
9], the total graph
is built on the full set of ring elements as vertices. Two distinct vertices
x and
y are connected if their sum
is a zero-divisor, i.e.,
. This summation-based adjacency condition introduces new layers of algebraic nuance, capturing not just multiplication-based annihilation, but additive interactions involving zero-divisors. Several subsequent investigations have examined total graphs over rings like
[
10], their complements [
11], and connectivity-related questions for more general finite rings [
12].
Further developments include analysis of regularity and edge conditions [
13], adaptations to lattice structures [
14], complements over finite fields [
15], and exploration of spectral properties such as eigenvalue bounds and multiplicities [
16]. These studies reflect the growing interest in combining algebraic insight with structural and analytic properties of the associated graphs.
The Cartesian product of rings presents a natural yet nontrivial extension for these graphical models. In particular, for direct products such as
, Dhorajia [
17] computed fundamental invariants of the total graph, including its connectivity, diameter, clique number, and independence number. Axtell et al. [
18] provided a similar analysis in the context of zero-divisor graphs over direct products. Despite such progress, research involving total graphs on larger Cartesian products, specifically products involving three or more rings, remains sparse and open to generalization.
The recent surge of interest in spectral aspects of algebraic graphs has also permeated the study of total graphs. Investigations involving normalized Laplacian and Randić spectra [
6,
7], alongside distance-based spectral measures [
8], highlight the relevance of these spectral invariants in distinguishing graph structures that originate from different algebraic sources. These findings reinforce the necessity for studying broader classes of rings and their associated graphs, especially those defined over higher-dimensional Cartesian products.
Purpose and Scope: This article extends the current framework by initiating a comprehensive study of the total graph over the three-fold Cartesian product of finite commutative rings:
We provide full derivations and formal proofs for a wide range of graph-theoretic properties including structural characteristics (connectivity, regularity, clique number, independence number), topological invariants (planarity and genus), and traversal features (Hamiltonicity, Eulerian properties). We also include an in-depth study of the automorphism group of the total graph and investigate its behavior in symmetric settings where .
Unlike previous two-component studies, the three-factor case introduces new challenges in combinatorial configurations, additive dependencies across three independent ring components, and increased genus growth, which cannot be easily deduced from binary cases. Moreover, symmetric cases yield nontrivial automorphism groups that necessitate group-theoretic tools such as semidirect products. These intricacies demand new constructions and sharper proof techniques, which we present throughout this manuscript.
Organization of the Paper:
Section 2 presents the necessary background, notation, and prior results needed for our analysis.
Section 3 outlines the methodology adopted to investigate the total graph
.
Section 4 contains the main contributions of the paper, structured into four subsections: structural, topological, and traversal properties, followed by automorphism group analysis. Finally,
Section 5 concludes with a summary of the results, proposed applications, and directions for future work.
3. Methodology
The present study extends the framework of total graphs from two-factor to three-factor Cartesian products of finite commutative rings. Throughout, we consider the ring with , and analyze its total graph . Our methodology intertwines algebraic characterization of zero-divisors with classical graph-theoretic techniques, thereby yielding rigorous and self-contained proofs of every stated result.
First, we formalise adjacency in . Vertices are the elements of R, and two distinct vertices and are adjacent exactly when their component-wise sum lies in the zero-divisor set of R. The description of follows directly from Lemma 1. This additive rule reduces many graph-theoretic questions to arithmetic properties of .
Building on Anderson and Badawi’s connectivity result for arbitrary finite rings [
9] and on Dhorajia’s two-factor analysis [
17], we lift structural parameters from the component graphs
to
. In particular, Theorems 1 and 2 ensure that
is connected with diameter at most 2, and that its regularity hinges on the parity of
. Clique and independence numbers are obtained via maximum– and minimum–type formulas, generalizing Lemma 2 to three components.
Topological features are addressed next. Using Euler’s formula together with the lower bound
for planar embeddings, we prove that
is non-planar whenever
and establish an improved genus bound. These arguments refine earlier genus estimates for zero-divisor graphs [
5] and adapt them to the total graph context.
Traversal properties are analyzed through classical criteria: Dirac’s theorem yields Hamiltonicity when at least one modulus is even, while the Eulerian condition reduces to the parity of vertex degrees—hence the simultaneous oddness of
. Additional investigations of bipartiteness and cycle structure reveal how the ring components influence graph traversal, complementing connectivity results of Lucas [
3].
Finally, we characterise the automorphism group
. By exploiting the independence of coordinates, we show that it decomposes into the direct product of the automorphism groups of the three component graphs. When
, an extra symmetry corresponding to permutations of the factors appears; this is formalised via a semidirect product, echoing the symmetry considerations of Chelvam and Asir [
10].
All theoretical findings are illustrated by explicit numerical Examples, chosen to validate the derived formulas and to demonstrate practical computation. Calculations were verified in SageMath; code is available on request. Collectively, these steps provide a comprehensive, reviewer-requested treatment of the structural, topological, and traversal properties of and lay groundwork for future extensions to higher-dimensional products.
Novelty and Generalization Strategy. Many of the results in this work extend known properties of total graphs over two-factor Cartesian products, such as
, to the more complex setting of three-factor products
. This generalization introduces significant combinatorial challenges, especially in proofs involving adjacency conditions, structural parameters, and symmetry arguments. Throughout
Section 4, we emphasize how these properties evolve in the presence of an additional ring component and demonstrate that key invariants—such as regularity, clique number, and automorphism group—reveal richer behavior not captured by the two-factor case.
4. Main Results
4.1. Structural Properties of
Theorem 3. Let , where . Then, the total graph is connected, and its diameter satisfies .
Proof. Let and be two distinct vertices in . We aim to show that either u and v are adjacent, or there exists a third vertex w such that and .
Case 1: . Then, by the definition of the total graph, u and v are directly adjacent, and .
Case 2: . Then, the coordinate-wise sum
lies outside
, which implies that all three components are regular:
Since
, we can construct a vertex
where exactly one coordinate is a zero-divisor and the other two are zero. For example, let us take:
Then
by definition, since the first coordinate is a zero-divisor.
Now compute:
Because
is a zero-divisor in
, at least one of
or
is a zero-divisor as well. Even if not, we can adjust
w to ensure adjacency.
In general, define
with:
Such a
w always exists since each component ring has at least one zero-divisor (as
). Then, at least one of the sums
and
must belong to
, ensuring
and
.
Thus, any two vertices in
are either directly connected or have a common neighbor, and so:
and the graph is connected. □
Example 1. Let . The zero-divisor sets of the component rings are:Hence,Consider the vertices and . Their sum is . Sincewe observe that , and thus u and v are not adjacent. Now consider the vertex , since:Computing the sums:we find thatso , and similarly,so . Therefore, both and hold, yielding a path of length 2 in . This confirms that the distance between u and v is at most 2, thereby illustrating Theorem 3.
Remark 1. The intermediate vertex constructed in the proof of Theorem 3 is always valid because the sum of any two distinct vertices in R necessarily differs in at least one coordinate. Since each , , and contains non-zero zero-divisors when , the existence of a component-wise zero-divisor guarantees that adjacency can be achieved through such a constructed intermediate.
Theorem 4. Let with . Then, the total graph is regular if and only if each of the moduli is even.
Proof. Recall that a graph is regular if all vertices have the same degree. Let
, and define its degree by
This count depends on the structure of the zero-divisor set
, which satisfies
Sufficiency. Assume
are all even. Then,
,
, and
. For any
, we have
. Since the translation
is a bijection of
R onto itself, and since
for all
v, it follows that
uniformly across all vertices. Hence
is regular.
Necessity. Suppose at least one modulus is odd; without loss of generality, let
n be odd. Then,
, so for any
,
. Consider the elements
We observe:
so
while
. Therefore,
The degrees differ, violating regularity. Thus, if
is regular, each of
must be even. □
Example 2. (a) Regular Case. Consider , where all moduli are even. The sets of zero-divisors are:By Theorem 4, is regular, and every vertex has a degree equal to . This follows since for all , and the additive structure guarantees uniformity in the neighborhood sizes. (b) Non-Regular Case. Let , where the modulus 5 is odd. Consider and . Then, , but . Hence,showing is not regular. Theorem 5. Let . Then, the clique number of the total graph satisfies Proof. Let for , and set .
Lower bound. Assume without loss of generality that
, and let
be a maximum clique in
of size
. Fix arbitrary
and
, and define
Since adjacency in
requires that the sum of any two vertices lies in
, and the zero-divisor condition is preserved in the first component for elements of
, it follows that
C is a clique in
of size
. Thus,
Upper bound. Let be any clique in , and write each vertex as . For every unordered pair , the adjacency condition ensures that , i.e., at least one component of the sum lies in the respective zero-divisor set. Color each edge according to which coordinate witnesses the zero-divisor condition (i.e., color 1 if , etc.). By the pigeonhole principle, at least edges share a color; assume color 1 achieves this maximum. Then, the projection of C onto must form a clique in , hence .
Combining both bounds yields the equality , as claimed. □
Example 3. Consider . Based on known results (see [10] (Thm. 3.1)), we have:Thus, Theorem 5 gives: A maximal clique of size 4 in is realized by fixing the second and third coordinates and selecting zero-divisors in the first component:Indeed, since and fixed values in the other coordinates preserve adjacency, all pairwise sums lie in . Theorem 6. Let . Then, the independence number of the total graph satisfies Proof. Let for , and set .
Upper bound. Let be an independent set in . For any two distinct elements , the condition must hold. This implies that in each coordinate projection, the image of I must form an independent set in the corresponding component graph. Hence, for each i, and so . Therefore, .
Lower bound. Assume without loss of generality that
. Let
be an independent set of maximum size
in
, and fix arbitrary elements
,
. Then, the set
is independent in
, since any two elements differ only in their second coordinate, and the sum of any two such coordinates avoids
. Hence,
, and
.
Combining both bounds yields the equality. □
Example 4. Let . Using results from [10] (Table 2), we find:Thus, Theorem 6 implies An explicit independent set of size 2 issince , and the other components are fixed. This confirms the bound is tight. 4.2. Topological Properties of
Theorem 7. Let with . Then, the total graph is non-planar.
Proof. Write
. For each vertex
, we estimate its degree. Choose fixed zero-divisors
,
,
(which exist because
). Form the following mutually distinct neighbours of
v:
Each differs from
v in at most two coordinates and their coordinate-wise sums with
v contain a zero-divisor, hence they are adjacent to
v. Thus, every vertex has a degree of at least 6; denote
.
The total number of edges is, therefore, Euler’s inequality for planar graphs ( when ) is violated, so cannot be planar. □
Remark 2. If , planarity depends on the detailed distribution of zero-divisors and requires separate analysis.
Example 5. Let . Here and every vertex degree is at least 6 by the construction in the proof, giving . Hence is non-planar, confirming Theorem 7.
This result strengthens and extends earlier genus bounds for two-factor total graphs by deriving a sharp lower bound for three-component Cartesian products, which significantly increases the embedding complexity.
Theorem 8. Let and denoteFor the total graph , one has the lower bound Proof. Embed
on an orientable surface of genus
g and let
F be the number of faces. Euler’s formula gives
. Since every face is bounded by at least three edges,
, i.e.,
. Substituting this maximum value for
F yields
so
and, therefore,
□
Remark 3. When one or more of is small, the bound in Theorem 8 may not be tight; direct computation or exhaustive embedding methods can then determine the exact genus.
Example 6. Consider . The vertex count is . From Example 5 each vertex has a degree of at least 12, giving Applying Theorem 8,Thus requires a surface of genus at least 106, far beyond the planar or toroidal cases. This result extends classical connectivity principles to the total graph of a three-factor ring product, showing that edge-connectivity remains tightly bound to the minimum degree, as previously observed in simpler ring constructions.
Theorem 9. Let . Let denote the edge-connectivity and the minimum vertex degree in the total graph. Then: Proof. Upper bound. Removing all edges incident to a vertex of minimum degree isolates that vertex from the graph, so we have:
Lower bound. By Theorem 3, the diameter of
satisfies
. A classical result due to Plesník [
19] (Theorem 1) asserts that for any finite simple connected graph with a diameter of at most 2, the edge-connectivity equals the minimum vertex degree. Applying this to
yields:
Combining the upper and lower bounds proves the equality. □
Example 7. Let . Each vertex has a degree of 10, so the minimum degree is . By Theorem 9, it follows that the edge-connectivity is also 10. Thus, removing any 10 edges incident to a single vertex disconnects it from the rest of the graph, while any smaller set of edge deletions leaves the graph connected.
This result unifies vertex- and edge-connectivity for three-component total graphs, extending prior results on two-factor rings and illustrating the tight interplay between structural robustness and minimum degree.
Theorem 10. Let . Then, for the total graph , the following identity holds:where κ is the vertex-connectivity, the edge-connectivity, and δ the minimum vertex degree of . Proof. From Theorem 3, the total graph
is connected and satisfies
. A classical result by Plesník [
19] (Theorem 1) states that for any simple connected graph with a diameter of at most 2,
Since
meets the required conditions, the equalities follow directly. □
Example 8. Let . A direct computation shows that the minimum degree of is . Thus, by Theorems 9 and 10, we have:This means that removing any 14 vertices keeps the graph connected, while removing a specific set of 15 vertices can disconnect it. 4.3. Traversal Properties of
In this section, we investigate how the graph behaves from a traversal viewpoint. Our focus lies on Hamiltonicity, Eulerian conditions, bipartiteness, and the structure of short cycles.
This result extends the classical Hamiltonicity condition to total graphs over three-fold Cartesian products of finite rings, using Dirac’s theorem to highlight the effect of parity on traversal feasibility.
Theorem 11. Let with . If at least one of is even, then the total graph is Hamiltonian.
Proof. Let
, and for any
k, let
denote Euler’s totient function. A vertex
is non-adjacent to a vertex
if and only if
i.e., all component-wise sums are regular. The total number of such
u equals
, which gives:
This expression is valid for every
, so the graph is regular and has minimum degree:
Assume without loss of generality that
n is even. Then, at least half the elements of
are even and hence zero-divisors, so
. For
, one has
, with equality for
. Therefore,
Then, the minimum degree satisfies
Since
implies
, the condition
holds. By Dirac’s theorem [
20] (Thm. 10.2), any simple graph with
and minimum degree
is Hamiltonian. Therefore,
contains a Hamiltonian cycle. □
Example 9. Let . Then , and Euler’s totient values are:Thus , soHence, by Theorem 11, the graph is Hamiltonian. Remark 4. If are all odd, the minimum degree δ may not satisfy Dirac’s condition. In such cases, Hamiltonicity depends on deeper arithmetic relationships between the ring components and requires further investigation. This remains an open direction for future work.
This result generalizes the Eulerian characterization from two-factor total graphs to the case of triple products, with parity conditions on Euler’s function governing degree constraints across all vertices.
Theorem 12. The total graph is Eulerian if and only if are all odd.
Proof. A finite graph is Eulerian if and only if it is connected and all vertex degrees are even. The connectivity of is guaranteed by Theorem 3, so it remains to analyze vertex degrees.
Let
, and fix a vertex
. A vertex
is
not adjacent to
v precisely when
Since
contains
elements, the number of such non-neighbors is
. Thus, the degree of
v is
so the parity of
matches that of
.
(Only-if part). Suppose at least one of is even. Then the corresponding totient is even (since if k is odd, and of any even number is even). Since the remaining two values are odd (if m and p are odd), the product is even. Then is odd for every v, contradicting the Eulerian condition.
(If part). Conversely, if are all odd, then each is even for , and so their product is even. Hence, is for all v, and satisfies the Eulerian criterion. □
Example 10. Let . Since , , and are all odd, Theorem 12 applies and confirms that is Eulerian.
We compute:Since all vertices have even degree and the graph is connected, it is indeed Eulerian. Remark 5. If any of is even, then is not Eulerian. However, it may still contain Eulerian subgraphs or Eulerian circuits within induced subgraphs, particularly in cases with structured parity distributions.
This theorem refines the bipartiteness characterization from previous two-factor studies by identifying disjointness conditions on non-zero zero-divisor sets in each component, ensuring the absence of odd cycles in three-fold total graphs.
Theorem 13. Let and define analogously. Then, the total graphis bipartite if and only if the three sets , , and are pairwise disjoint. Proof. (If part). Assume that
. Define a bipartition
of the vertex set
as follows:
Because
by definition, the zero vector
lies in
B, ensuring
B is nonempty.
Suppose . Then, for some coordinate position, say the first, we have either or , or similarly for the second or third components. Since the D-sets are pairwise disjoint, the sum cannot fall into —because none of the coordinate sums simultaneously hit a zero-divisor in more than one component. Hence, u and v are not adjacent in , implying A is an independent set; the same applies to B. Thus is bipartite.
(Only-if part). Suppose instead that
, and let
. Then, consider the vertices:
Observe:
Since
and
, and the sum of
d with itself stays in
(due to closure under addition modulo
n), all these pairs are adjacent, forming a 3-cycle. Thus
is not bipartite. An identical construction holds if
or
. Therefore, bipartiteness requires that
be pairwise disjoint. □
Example 11. Let . We compute:Here , violating the disjointness condition. Hence by Theorem 13, is not bipartite. Indeed, the verticesform a 3-cycle since each pair adds to an element in , confirming non-bipartiteness. This result provides the first precise girth characterization for total graphs over three-component rings of 2-power modulus, confirming bipartiteness and minimal cycle length based on zero-divisor structure.
Theorem 14. Let , , and with . Then the total graph - (a)
is bipartite (and thus contains no odd cycles), and
- (b)
contains an induced 4-cycle.
Consequently, the girth of the graph equals 4.
Proof. (a) Bipartiteness. For any ring , the non-zero zero-divisors form the set of even residues . Because are distinct powers of 2, the corresponding D-sets are disjoint. By Theorem 13, the graph is bipartite, and hence, contains no odd cycles.
(b) Existence of a 4-cycle. Let
Define the four vertices:
We observe:
while
and
are not in
since no component sum alone hits a zero-divisor.
Hence, the vertices form an induced 4-cycle. Since the graph is bipartite and admits a 4-cycle, it follows that the girth is exactly 4. □
Example 12. Let . Take , , and . Then, the verticesform a 4-cycle as constructed in the proof. Moreover, since all moduli are powers of 2, the graph is bipartite and hence has no odd-length cycles. Therefore, the girth of is 4. 4.4. Automorphism Group
We now analyze the automorphism group of the total graph . For a graph G, the automorphism group comprises all bijective mappings from to preserve the adjacency itself.
This result generalizes known results on automorphism groups of total graphs over two-factor products by describing the full automorphism structure of three-factor total graphs as a direct product of component groups. In symmetric cases, additional permutations arise, yielding a semidirect product.
Theorem 15. Let with . Then, the automorphism group of its total graph satisfies Proof. Let
and denote the vertex set by
. Recall that adjacency is defined by:
Any automorphism
must preserve adjacency, and hence must act independently on each coordinate to preserve the zero-divisor conditions. This induces a triple of automorphisms:
Conversely, any triple of automorphisms
from the component rings defines a graph automorphism of
G via:
This correspondence is bijective and operation-preserving, and thus defines an isomorphism:
□
Example 13. Let . From [10] and direct computations, we have:Applying Theorem 15, we conclude: Remark 6. When , an additional layer of symmetry arises due to the indistinguishability of coordinates. In this case, the automorphism group expands to include all permutations of the three components:where acts by permuting the factors, and the semidirect product accounts for interactions between coordinate permutations and local automorphisms. Theorem 16. When all three ring components are identical, the total graph admits enhanced symmetry via coordinate permutations. Let with . Then,where acts by permuting the three coordinates. Proof. By Theorem 15, every automorphism of
arises from a triple of component-wise automorphisms
Since all three components are isomorphic, any permutation
of the coordinates also preserves adjacency and defines an automorphism of
via
Combining these two symmetries, define the map
which establishes a group homomorphism from the semidirect product
This map is bijective and respects the group operation, hence it is an isomorphism. □
Example 14. Let . From [10], the total graph has automorphism group isomorphic to the symmetric group . Applying Theorem 16, we obtainHere the normal subgroup corresponds to independent actions on each coordinate, while the semidirect component encodes permutations of the three factors. This result gives a complete characterization of the automorphism group of the total graph over a prime field, identifying it as a wreath product based on symmetric and involutive symmetries.
Theorem 17. Let p be an odd prime. Write and denoteThen, consists of r independent edges joining the pairs in and an isolated vertex 0. Its automorphism group isthat is, the wreath product acting by independent swaps inside each pair and permutations of the r pairs. For , is edgeless on two vertices and . Proof. Since , two distinct elements are adjacent in if and only if , i.e., . The non-zero elements thus form r independent edges of the form for , and the vertex 0 is isolated.
Each such edge admits a swap symmetry, giving a factor isomorphic to . Any permutation of the r edge pairs also preserves adjacency, yielding an factor. Together these symmetries generate all automorphisms of , giving the wreath product . For , has two vertices , no edges, and full symmetric group as its automorphism group. □
Example 15. Let so that . The non-zero units modulo 7 pair up asHence the graph consists of the three disjoint edges , , and the isolated vertex 0. The automorphism group is the wreath product , capturing all edge swaps and pair permutations—exactly as described in Theorem 17. This result demonstrates that for rings of the form , the automorphism group of the total graph acts at least as the full symmetric group on its non-zero zero-divisors, revealing rich internal symmetry.
Theorem 18. Let with , and defineThe restriction mapis surjective. In particular, Proof. For with , all non-zero even elements of are zero-divisors, forming the set . Since the sum of any two even elements is again even, the induced subgraph on is complete. Moreover, the element 0 is adjacent to every vertex in , while the unique odd unit 1 is not adjacent to any vertex in .
Now, let be an arbitrary permutation of . Extend to a bijection on by fixing 0 and 1, and defining . Since adjacency is preserved within and 0 remains adjacent to every element of , while 1 retains its non-neighbourhood, is an automorphism of . This proves surjectivity of the restriction map and confirms the subgroup containment. □
Example 16. For , the set of non-zero zero-divisors isThe induced subgraph on is the complete graph , and 0 is adjacent to all three. By Theorem 18, we haveIndeed, any permutation of that fixes 0 and 1 defines an automorphism of the total graph. This result fully generalizes the structure of the automorphism group of the total graph over multi-factor finite rings, capturing both coordinate-wise symmetry and permutation symmetry through a semidirect product.
Theorem 19. Let with . Then, the automorphism group of the total graph satisfieswhere acts on the product by permuting the k coordinate factors. Proof. Since adjacency in is determined by coordinate-wise addition and membership in the zero-divisor sets of the respective rings, any automorphism must preserve these structural constraints.
Every automorphism
therefore restricts to a tuple of coordinate-wise automorphisms
and may be accompanied by a permutation
acting on the coordinate indices. This defines a natural semidirect product structure.
Conversely, given a tuple of automorphisms
and a permutation
, define a map
This bijection preserves adjacency in
due to the component-wise definition of adjacency and the fact that permutations of identical components retain algebraic properties. The map defines a valid automorphism, completing the isomorphism. □
Example 17. Let (). Since and are edgeless graphs on two and three vertices, respectively, one haswhile by [10]. Applying Theorem 19, we obtainHere acts by permuting the three coordinate rings; however, it acts nontrivially only on the third factor via conjugation, since the first two automorphism groups are trivial. 5. Conclusions and Future Work
This paper generalizes the structural and combinatorial study of total graphs from two-factor to three-factor Cartesian products of finite commutative rings. For the total graph we presented a detailed investigation of its connectivity, regularity, topological properties, traversal features, and symmetry behavior.
From a structural perspective, we proved that
is always connected with diameter at most two, and regular if and only if all moduli are even. The clique and independence numbers are governed by extremal values among those of the component total graphs. Topologically, we demonstrated non-planarity when
and established a sharp lower bound on the genus,
which shows rapid topological growth as ring sizes increase.
Traversal properties were resolved using classical theorems: the graph is Hamiltonian under mild conditions (e.g., one modulus even) and Eulerian precisely when all moduli are odd. Bipartiteness is fully characterized by a disjointness condition on zero-divisor subsets, while girth was shown to be exactly 4 when the moduli are powers of two.
In terms of graph symmetry, we proved that the automorphism group of decomposes as the direct product of the automorphism groups of the component total graphs. When the moduli coincide, an additional symmetry appears due to coordinate permutations, resulting in a semidirect product structure. We further extended this to a general semidirect product formula for k-fold Cartesian products, thereby unifying the behavior of symmetries in all such total graphs.
In summary, this work offers a foundational study of total graphs over three-ring Cartesian products, bridging algebra and graph theory in a novel way. The results presented herein both generalize prior work and lay the groundwork for future developments across structural graph theory, algebraic combinatorics, and applied network design.