
Citation: Chris Cadonic, Benedict C. Albensi. Oscillations and NMDA Receptors: Their Interplay Create Memories[J]. AIMS Neuroscience, 2014, 1(1): 52-64. doi: 10.3934/Neuroscience.2014.1.52
[1] | Jingli Xie, Hongli Guo, Meiyang Zhang . Dynamics of an SEIR model with media coverage mediated nonlinear infectious force. Mathematical Biosciences and Engineering, 2023, 20(8): 14616-14633. doi: 10.3934/mbe.2023654 |
[2] | Martin Luther Mann Manyombe, Joseph Mbang, Jean Lubuma, Berge Tsanou . Global dynamics of a vaccination model for infectious diseases with asymptomatic carriers. Mathematical Biosciences and Engineering, 2016, 13(4): 813-840. doi: 10.3934/mbe.2016019 |
[3] | Hamdy M. Youssef, Najat A. Alghamdi, Magdy A. Ezzat, Alaa A. El-Bary, Ahmed M. Shawky . A new dynamical modeling SEIR with global analysis applied to the real data of spreading COVID-19 in Saudi Arabia. Mathematical Biosciences and Engineering, 2020, 17(6): 7018-7044. doi: 10.3934/mbe.2020362 |
[4] | Abdisa Shiferaw Melese, Oluwole Daniel Makinde, Legesse Lemecha Obsu . Mathematical modelling and analysis of coffee berry disease dynamics on a coffee farm. Mathematical Biosciences and Engineering, 2022, 19(7): 7349-7373. doi: 10.3934/mbe.2022347 |
[5] | Yujie Sheng, Jing-An Cui, Songbai Guo . The modeling and analysis of the COVID-19 pandemic with vaccination and isolation: a case study of Italy. Mathematical Biosciences and Engineering, 2023, 20(3): 5966-5992. doi: 10.3934/mbe.2023258 |
[6] | María Guadalupe Vázquez-Peña, Cruz Vargas-De-León, Jorge Fernando Camacho-Pérez, Jorge Velázquez-Castro . Analysis and Bayesian estimation of a model for Chikungunya dynamics with relapse: An outbreak in Acapulco, Mexico. Mathematical Biosciences and Engineering, 2023, 20(10): 18123-18145. doi: 10.3934/mbe.2023805 |
[7] | S. J. Gutowska, K. A. Hoffman, K. F. Gurski . The effect of PrEP uptake and adherence on the spread of HIV in the presence of casual and long-term partnerships. Mathematical Biosciences and Engineering, 2022, 19(12): 11903-11934. doi: 10.3934/mbe.2022555 |
[8] | Maryam Al-Yahyai, Fatma Al-Musalhi, Ibrahim Elmojtaba, Nasser Al-Salti . Mathematical analysis of a COVID-19 model with different types of quarantine and isolation. Mathematical Biosciences and Engineering, 2023, 20(1): 1344-1375. doi: 10.3934/mbe.2023061 |
[9] | Yiwei He, Qianqian Cui, Zengyun Hu . Modeling and analysis of the transmission dynamics of cystic echinococcosis: Effects of increasing the number of sheep. Mathematical Biosciences and Engineering, 2023, 20(8): 14596-14615. doi: 10.3934/mbe.2023653 |
[10] | Rajanish Kumar Rai, Pankaj Kumar Tiwari, Yun Kang, Arvind Kumar Misra . Modeling the effect of literacy and social media advertisements on the dynamics of infectious diseases. Mathematical Biosciences and Engineering, 2020, 17(5): 5812-5848. doi: 10.3934/mbe.2020311 |
By a graph $ G = (V(G), E(G)) $, with $ V(G) $ and $ E(G) $ the set of vertices and edges of $ G $ respectively, we mean an undirected simple graph without isolated vertices (i.e., each vertex has at least a neighbor).
Given a graph $ G $, representing a chemical structure,
$ X(G) = \sum\limits_{uv \in E(G)} F(d_u, d_v) $ |
is said a topological descriptor and, if also it correlates with a molecular property, it is called a topological index. Above, by $ uv $ we mean the edge of a graph $ G $ joining the vertices $ u $ and $ v $, $ d_u $ denotes the degree of $ u $, and $ F $ is an appropriate chosen function. Remarkably, topological indices capture physical properties of a chemical compound in a single number.
A great number of topological indices have been defined and studied over more than four decades. Among them, probably the most popular topological indices are the Randić and the Zagreb indices. The first and second Zagreb indices, denoted by $ M_1 $ and $ M_2 $, respectively, were defined by Gutman and Trinajstić (see [1]) in $ 1972 $ by
$ M_1(G) = \sum\limits_{u\in V(G)} d_u^2, \qquad M_2(G) = \sum\limits_{uv\in E(G)} d_u d_v . \qquad $ |
For more details of the applications and mathematical properties of Zagreb indices see [2,3,4], and the references therein. Zagreb indices have many connections with other topological indices, see e.g., [5,6].
The concept of variable molecular descriptors was proposed as a way of characterizing heteroatoms (see [7,8]), but also to assess structural differences in alkylcycloalkanes [9]. The idea behind the variable molecular descriptors is that the variables are determined during the regression in order to minimize the error of estimate for a particular chemical property (see, e.g., [10]).
In this line of ideas, the variable versions of the first and second Zagreb indices were introduced as [10,11,12]
$ M_1^{ \alpha}(G) = \sum\limits_{u\in V(G)} d_u^{ \alpha}, \qquad M_2^{ \alpha}(G) = \sum\limits_{uv\in E(G)} (d_u d_v)^ \alpha , $ |
with $ \alpha \in \mathbb{R} $. Evidently, $ M_1^{2} $ and $ M_2^{1} $ are the first and second Zagreb indices, respectively. In addition, the first and second variable Zagreb indices include several known indices. As examples we note that $ M_2^{-1/2} $ is the Randić index, $ M_1^{3} $ is the forgotten index $ F $, $ M_1^{-1} $ is the inverse index $ ID $, and $ M_2^{-1} $ is the modified Zagreb index.
In 2011, Vukičević proposed the variable symmetric division deg index [13]
$ SDDα(G)=∑uv∈E(G)(dαudαv+dαvdαu). $
|
(1.1) |
Note that $ SDD_{-\alpha}(G) = SDD_\alpha(G) $ and so, it suffices to consider positive values of $ \alpha $. The symmetric division deg index is the best predictor of total surface area for polychlorobiphenyls [14].
In this work we perform studies of the variable symmetric division deg index from analytical and computational viewpoints. We obtain new inequalities for the variable symmetric division deg index $ SDD_\alpha(G) $ and we characterize graphs extremal with respect to them. Some of these inequalities generalize and improve previous results for the symmetric division deg index. In addition, we computationally apply the $ SDD_\alpha(G) $ index on random graphs and we demonstrate that the ratio $ \langle SDD_\alpha(G) \rangle/n $ ($ n $ denotes the order of the graph) depends only on the average degree $ \langle d \rangle $.
One of our main results is Theorem 8, which provides upper and lower bounds of $ SDD_\alpha(G) $ in terms of the number of edges, the maximum and the minimum degree of $ G $.
Let us start by proving a monotonicity property of these indices.
Theorem 1. Let $ G $ be a graph and $ 0 < \alpha < \beta $.Then
$ SDD_{ \alpha}(G) \le SDD_{ \beta}(G) , $ |
and the equality in the bound is attained if and only if each connected component of $ G $ is a regular graph.
Proof. Let us consider $ x \ge 1 $. Thus, $ x^{ \alpha} \ge x^{- \beta} $ and
$ xβ−α−1≥0,xα(xβ−α−1)≥x−β(xβ−α−1),xβ−xα≥x−α−x−β,xβ+x−β≥xα+x−α, $
|
for every $ x \ge 1 $. Since $ u(x) = x^{ \alpha} + x^{- \alpha} $ satisfies $ u(1/x) = u(x) $ for every $ x > 0 $, we have $ x^{ \beta} + x^{- \beta} \ge x^{ \alpha} + x^{- \alpha} $ for every $ x > 0 $. Note that we obtain the equality if and only if $ x = 1 $.
Thus, we have
$ SDDβ(G)=∑uv∈E(G)(dβudβv+dβvdβu)≥∑uv∈E(G)(dαudαv+dαvdαu)=SDDα(G). $
|
The previous argument gives that we have the equality in the bound if and only if $ d_u/d_v = 1 $ for every $ uv \in E(G) $, i.e., each connected component of $ G $ is a regular graph.
Our next results in this section provide bounds of $ SDD_\alpha(G) $ involving the maximum and minimum degree of the graph $ G $. Since scientists often estimate average degree of large networks, we present in the next section results involving the average degree.
Our next theorem relates the $ SDD_ \alpha $ and the variable Zagreb indices.
Theorem 2. If $ G $ is a graph with minimum degree $ \delta $ and maximum degree $ \Delta $, and $ \alpha > 0 $, then
$ 2δ2αM−α2(G)≤SDDα(G)≤2Δ2αM−α2(G),Δ−2αM2α+11(G)≤SDDα(G)≤δ−2αM2α+11(G), $
|
and we have the equality in each bound if and only if $ G $ is regular.
Proof. First of all recall that for every function $ f $ the following equality
$ \sum\limits_{uv \in E(G)} \big( f(d_u)+ f(d_v) \big) = \sum\limits_{u \in V(G)} d_u f(d_u) $ |
holds. In particular,
$ \sum\limits_{uv \in E(G)} \big( d_u^{2 \alpha}+d_v^{2 \alpha} \big) = \sum\limits_{u \in V(G)} d_u^{2 \alpha+1} = M_1^{2a+1}(G). $ |
Since
$ SDD_{ \alpha}(G) = \sum\limits_{uv \in E(G)} \!\! \left( \, \frac{d_u^{ \alpha}}{d_v^{ \alpha}} + \frac{d_v^{ \alpha}}{d_u^{ \alpha}} \right) = \sum\limits_{uv \in E(G)} \frac{d_u^{2 \alpha}+d_v^{2 \alpha}}{(d_u d_v)^{ \alpha}} \, . $ |
and $ \alpha > 0 $, we obtain
$ SDD_{ \alpha}(G) = \sum\limits_{uv \in E(G)} \frac{d_u^{2 \alpha}+d_v^{2 \alpha}}{(d_u d_v)^{ \alpha}} \le 2 \Delta^{2 \alpha} \!\! \!\! \sum\limits_{uv \in E(G)} (d_u d_v)^{- \alpha} = 2 \Delta^{2 \alpha} M_2^{- \alpha}(G) , $ |
and
$ SDD_{ \alpha}(G) = \sum\limits_{uv \in E(G)} \frac{d_u^{2 \alpha}+d_v^{2 \alpha}}{(d_u d_v)^{ \alpha}} \ge 2 \delta^{2 \alpha} \!\! \!\! \sum\limits_{uv \in E(G)} (d_u d_v)^{- \alpha} = 2 \delta^{2 \alpha} M_2^{- \alpha}(G) . $ |
We also have
$ SDDα(G)=∑uv∈E(G)d2αu+d2αv(dudv)α≤δ−2α∑uv∈E(G)(d2αu+d2αv)=δ−2α∑u∈V(G)d2α+1u=δ−2αM2α+11(G), $
|
and
$ SDDα(G)=∑uv∈E(G)d2αu+d2αv(dudv)α≥Δ−2α∑uv∈E(G)(d2αu+d2αv)=Δ−2α∑u∈V(G)d2α+1u=Δ−2αM2α+11(G). $
|
If $ G $ is a regular graph, then each lower bound and its corresponding upper bound are the same, and both are equal to $ SDD_{ \alpha}(G) $.
Assume now that the equality in either the first or second bound holds. The previous argument gives that we have either $ d_u^{2 \alpha}+d_v^{2 \alpha} = 2 \Delta^{2 \alpha} $ for any $ uv \in E(G) $ or $ d_u^{2 \alpha}+d_v^{2 \alpha} = 2 \delta^{2 \alpha} $ for any $ uv \in E(G) $. Since $ \alpha > 0 $, we have $ \delta^{2 \alpha} \le d_u^{2 \alpha}, d_v^{2 \alpha} \le \Delta^{2 \alpha} $, and we conclude that $ d_u = d_v = \Delta $ for any $ uv\in E(G) $, or $ d_u = d_v = \delta $ for any $ uv\in E(G) $. Hence, $ G $ is a regular graph.
Finally, assume that the equality in either the third or fourth bound holds. The previous argument gives that we have either $ (d_u d_v)^{ \alpha} = \delta^{2 \alpha} $ for any $ uv \in E(G) $ or $ (d_u d_v)^{ \alpha} = \Delta^{2 \alpha} $ for any $ uv \in E(G) $. Since $ \alpha > 0 $, we have $ \delta^{ \alpha} \le d_u^{ \alpha}, d_v^{ \alpha} \le \Delta^{ \alpha} $, and we conclude that $ d_u = d_v = \delta $ for any $ uv\in E(G) $, or $ d_u = d_v = \Delta $ for every $ uv\in E(G) $. Therefore, $ G $ is a regular graph.
We will need the following technical result.
Lemma 3. Let $ 0 < a < A $. Then
$ a \le \frac{x^{2}+y^{2}}{x + y} \le A $ |
for every $ a \le x, y \le A $.The lower bound is attained if and only if $ x = y = a $.The upper bound is attained if and only if $ x = y = A $.
Proof. If $ a \le x, y \le A $, then $ ax+ay\le x^{2}+y^{2} \le Ax+Ay $, and the statement holds.
A large kind of topological indices, named Adriatic indices, was introduced in [14,15]. Twenty of them were selected as significant predictors of chemical properties. One of them, the inverse sum indeg index, defined by
$ ISI(G) = \sum\limits_{uv\in E(G)} \frac{d_u\, d_v}{d_u + d_v} = \sum\limits_{uv\in E(G)} \frac{1}{\frac{1}{d_u} + \frac{1}{d_v}}\, . $ |
appears in [14,15] as a good predictor of total surface area of octane isomers.
Next, we relate $ SDD_{ \alpha}(G) $ with the variable inverse sum deg index defined, for each $ a \in \mathbb{R} $, as
$ ISD_a(G) = \sum\limits_{uv \in E(G)} \frac{1}{d_u^a + d_v^a} \, . $ |
Note that $ ISD_{-1} $ is the inverse sum indeg index $ ISI $.
Theorem 4. If $ G $ is a graph with $ m $ edges and minimum degree $ \delta $, and $ \alpha > 0 $, then
$ SDD_{ \alpha}(G) \ge \frac{ \delta^{ \alpha} m^2}{ISD_{- \alpha}(G)} \, , $ |
and the equality in the bound holds if and only if $ G $ is regular.
Proof. Lemma 3 gives
$ \delta^{ \alpha} \le \frac{x^{2 \alpha}+y^{2 \alpha}}{x^{ \alpha} + y^{ \alpha}} \le \Delta^{ \alpha}, \qquad \frac1{x^{2 \alpha}+y^{2 \alpha}} \le \frac{ \delta^{- \alpha}}{x^{ \alpha} + y^{ \alpha}} \, , \qquad \frac1{d_u^{2 \alpha}+d_v^{2 \alpha}} \le \frac{ \delta^{- \alpha}}{d_u^{ \alpha} + d_v^{ \alpha}} \, , $ |
for every $ \delta \le x, y \le \Delta $. This last inequality and Cauchy-Schwarz inequality give
$ m2=(∑uv∈E(G)1)2=(∑uv∈E(G)(d2αu+d2αvdαudαv)1/2(dαudαvd2αu+d2αv)1/2)2≤∑uv∈E(G)d2αu+d2αvdαudαv∑uv∈E(G)dαudαvd2αu+d2αv≤δ−αSDDα(G)∑uv∈E(G)dαudαvdαu+dαv=δ−αSDDα(G)∑uv∈E(G)1d−αu+d−αv=δ−αSDDα(G)ISD−α(G). $
|
If $ G $ is a regular graph, then $ SDD_{ \alpha}(G) = 2m $, $ ISD_{- \alpha}(G) = m \delta^{ \alpha}/2 $ and the equality in the bound holds.
Assume now that the equality in the bound holds. Thus, by the previous argument,
$ \frac1{d_u^{2 \alpha}+d_v^{2 \alpha}} = \frac{ \delta^{- \alpha}}{d_u^{ \alpha} + d_v^{ \alpha}} $ |
for any $ uv \in E(G) $. Then Lemma 3 gives $ d_u = d_v = \delta $ for any $ uv\in E(G) $. Hence, $ G $ is a regular graph.
The modified Narumi-Katayama index is defined by
$ NK^*(G) = \prod\limits_{u\in V (G)} d_u^{d_u} = \prod\limits_{uv\in E (G)} d_u d_v $ |
in [16], inspired in the Narumi-Katayama index [17]. Next, we present an inequality relating $ SDD_{ \alpha}(G) $ and $ NK^*(G) $.
Theorem 5. Let $ G $ be a graph with $ m $ edges and minimum degree $ \delta $, and $ \alpha > 0 $.Then
$ SDD_ \alpha(G) \ge 2 \delta^{2 \alpha} m NK^*(G)^{- \alpha/m}, $ |
and the equality in the bound holds if and only if $ G $ is a regular graph.
Proof. Since the geometric mean is at most the arithmetic mean, we have
$ 1mSDDα(G)=1m∑uv∈E(G)(dαudαv+dαvdαu)=1m∑uv∈E(G)d2αu+d2αv(dudv)α≥2δ2α1m∑uv∈E(G)1(dudv)α≥2δ2α(∏uv∈E(G)1(dudv)α)1/m=2δ2αNK∗(G)−α/m. $
|
If $ G $ is a regular graph, then
$ 2 \delta^{2 \alpha} m NK^*(G)^{- \alpha/m} = 2 \delta^{2 \alpha} m \big( \delta^{2 m} \big)^{- \alpha/m} = 2m = SDD_ \alpha(G) . $ |
Finally, assume that the equality in the bound holds. The previous argument gives that $ d_u^{2 \alpha}+d_v^{2 \alpha} = 2 \delta^{2 \alpha} $ for any $ uv \in E(G) $. Since $ \alpha > 0 $, we obtain $ \delta^{2 \alpha} \le d_u^{2 \alpha}, d_v^{2 \alpha} $, and we have $ d_u = d_v = \delta $ for any $ uv \in E(G) $. Hence, $ G $ is regular.
Next, we obtain additional bounds of $ SDD_{ \alpha} $.
Theorem 6. Let $ G $ be a graph with $ m $ edges, minimum degree $ \delta $ and maximum degree $ \delta+1 $, $ \alpha > 0 $ and let $ A $ be the number of edges $ uv \in E(G) $ with $ d_u \ne d_v $.Then $ A $ is an even integer and
$ SDD_{ \alpha}(G) = 2m + A \left( \frac{( \delta+1)^{ \alpha}}{ \delta^{ \alpha}} + \frac{ \delta^{ \alpha}}{( \delta+1)^{ \alpha}} - 2 \right) . $ |
Proof. Let $ F = \left\lbrace uv\in E(G):\; d_u \ne d_v \right\rbrace $, then $ A $ is the cardinality of the set $ F $. Since the maximum degree of $ G $ is $ \delta+1 $ and its minimum degree is $ \delta $, if $ uv\in F $, then $ d_u = \delta $ and $ d_v = \delta+1 $ or viceversa, and therefore
$ \frac{d_u^{ \alpha}}{d_v^{ \alpha}} + \frac{d_v^{ \alpha}}{d_u^{ \alpha}} = \frac{( \delta+1)^{ \alpha}}{ \delta^{ \alpha}} + \frac{ \delta^{ \alpha}}{( \delta+1)^{ \alpha}} \, . $ |
If $ uv\in F^c = E(G) \setminus F $, then $ d_u = d_v = \delta $ or $ d_u = d_v = \delta+1 $, and therefore
$ \frac{d_u^{ \alpha}}{d_v^{ \alpha}} + \frac{d_v^{ \alpha}}{d_u^{ \alpha}} = 2 . $ |
Since there are exactly $ A $ edges in $ F $ and $ m-A $ edges in $ F^c $, we have
$ SDDα(G)=∑uv∈E(G)(dαudαv+dαvdαu)=∑uv∈Fc(dαudαv+dαvdαu)+∑uv∈F(dαudαv+dαvdαu)=∑uv∈Fc2+∑uv∈F((δ+1)αδα+δα(δ+1)α)=2m−2A+A((δ+1)αδα+δα(δ+1)α). $
|
This gives the equality.
Seeking for a contradiction assume that $ A $ is an odd integer.
Let $ \Gamma_1 $ be a subgraph of $ G $ obtained as follows: $ \Gamma_1 $ is induced by the $ n_1 $ vertices with degree $ \delta $ in $ V(G) $; denote by $ m_1 $ the number of edges of $ \Gamma_1 $. Handshaking Lemma implies $ n_1 \delta - A = 2m_1 $. Since $ A $ is an odd integer, $ \delta $ is also an odd integer. Thus, $ \delta+1 $ is even.
Let $ \Gamma_2 $ be the subgraph of $ G $ induced by the $ n_2 $ vertices with degree $ \delta+1 $ in $ V(G) $, and denote by $ m_2 $ the number of edges of $ \Gamma_2 $. Handshaking Lemma implies $ n_2(\delta+1) - A = 2m_2 $, a contradiction, since $ A $ is odd and $ \delta+1 $ is even.
Thus, we conclude that $ A $ is an even integer.
We will need the following result in the proof of Theorem 8 below.
Lemma 7. Given $ \alpha > 0 $, consider the function $ u: (0, \infty) \rightarrow (0, \infty) $ defined as $ u(t) = t^{ \alpha}+t^{- \alpha} $.Then $ u $ strictly decreases on $ (0, 1] $, $ u $ strictly increases on $ [1, \infty) $ and $ u(t) \ge u(1) = 2 $.
Proof. We have
$ u'(t) = \alpha t^{ \alpha-1}- \alpha t^{- \alpha-1} = \alpha t^{- \alpha-1}( t^{2 \alpha}-1). $ |
Since $ \alpha > 0 $, we have $ u' < 0 $ on $ (0, 1) $ and $ u' > 0 $ on $ (1, \infty) $. This gives the result. The following figure shows the function $ u(t) $ for some values of $ \alpha $.
Theorem 6 gives the precise value of $ SDD_{ \alpha} $ when $ \Delta = \delta+1 $. Theorem 8 below provides a lower bound when $ \Delta > \delta+1 $.
Theorem 8. Let $ G $ be a graph with $ m $ edges, minimum degree $ \delta $ and maximum degree $ \Delta > \delta+1 $. Denote by $ A_0, A_1, A_2, $ the cardinality of the subsets of edges $ F_0 = \{ uv \in E(G) : \, d_{u} = \delta, d_{v} = \Delta \} $, $ F_1 = \{ uv \in E(G) : \, d_{u} = \delta, \delta < d_{v} < \Delta \} $, $ F_2 = \{ uv \in E(G) : \, d_{u} = \Delta, \delta < d_{v} < \Delta \} $, respectively. If $ \alpha > 0 $, then
$ SDDα(G)≤(m−A1−A2)(Δαδα+δαΔα)+A1((Δ−1)αδα+δα(Δ−1)α)+A2(Δα(δ+1)α+(δ+1)αΔα),SDDα(G)≥2m+A0(Δαδα+δαΔα−2)+A1((δ+1)αδα+δα(δ+1)α−2)+A2(Δα(Δ−1)α+(Δ−1)αΔα−2). $
|
Proof. Lemma 7 gives that the function
$ \frac{d_{v}^{ \alpha}}{ \delta^{ \alpha}} + \frac{ \delta^{ \alpha}}{d_{v}^{ \alpha}} = u \left( \frac{d_{v}}{ \delta} \right) $ |
is increasing in $ d_{v} \in [\delta+1, \Delta-1] $ and so,
$ \frac{( \delta+1)^{ \alpha}}{ \delta^{ \alpha}} + \frac{ \delta^{ \alpha}}{( \delta+1)^{ \alpha}} \le \frac{d_{v}^{ \alpha}}{ \delta^{ \alpha}} + \frac{ \delta^{ \alpha}}{d_{v}^{ \alpha}} \le \frac{( \Delta-1)^{ \alpha}}{ \delta^{ \alpha}} + \frac{ \delta^{ \alpha}}{( \Delta-1)^{ \alpha}} \, , $ |
for every $ uv \in F_1 $.
In a similar way, Lemma 7 gives that the function
$ \frac{ \Delta^{ \alpha}}{d_{v}^{ \alpha}} + \frac{d_{v}^{ \alpha}}{ \Delta^{ \alpha}} = u \left( \frac{d_{v}}{ \Delta} \right) $ |
is decreasing in $ d_{v} \in [\delta+1, \Delta-1] $ and so,
$ \frac{ \Delta^{ \alpha}}{( \Delta-1)^{ \alpha}} + \frac{( \Delta-1)^{ \alpha}}{ \Delta^{ \alpha}} \le \frac{ \Delta^{ \alpha}}{d_{v}^{ \alpha}} + \frac{d_{v}^{ \alpha}}{ \Delta^{ \alpha}} \le \frac{ \Delta^{ \alpha}}{( \delta+1)^{ \alpha}} + \frac{( \delta+1)^{ \alpha}}{ \Delta^{ \alpha}} \, , $ |
for every $ uv \in F_2 $.
Also,
$ 2 \le \frac{d_{u}^{ \alpha}}{d_{v}^{ \alpha}} + \frac{d_{v}^{ \alpha}}{d_{u}^{ \alpha}} \le \frac{ \Delta^{ \alpha}}{ \delta^{ \alpha}} + \frac{ \delta^{ \alpha}}{ \Delta^{ \alpha}} $ |
for any $ uv \in E(G) $.
We obtain
$ SDDα(G)=∑uv∈E(G)∖(F0∪F1∪F2)(dαudαv+dαvdαu)+∑uv∈F0(dαudαv+dαvdαu)+∑uv∈F1(dαudαv+dαvdαu)+∑uv∈F2(dαudαv+dαvdαu)≥∑uv∈E(G)∖(F0∪F1∪F2)2+∑uv∈F0(Δαδα+δαΔα)+∑uv∈F1(dαvδα+δαdαv)+∑uv∈F2(Δαdαv+dαvΔα). $
|
Hence,
$ SDDα(G)≥2m−2A0−2A1−2A2+A0(Δαδα+δαΔα)+A1((δ+1)αδα+δα(δ+1)α)+A2(Δα(Δ−1)α+(Δ−1)αΔα). $
|
We also have
$ SDDα(G)=∑uv∈E(G)∖(F1∪F2)(dαudαv+dαvdαu)+∑uv∈F1(dαudαv+dαvdαu)+∑uv∈F2(dαudαv+dαvdαu)≤(m−A1−A2)(Δαδα+δαΔα)+A1((Δ−1)αδα+δα(Δ−1)α)+A2(Δα(δ+1)α+(δ+1)αΔα). $
|
Here we deal with two classes of random graphs $ G $: Erdös-Rényi (ER) graphs $ G(n, p) $ and bipartite random (BR) graphs $ G(n_1, n_2, p) $. ER graphs are formed by $ n $ vertices connected independently with probability $ p \in [0, 1] $. While BR graphs are composed by two disjoint sets, sets 1 and 2, with $ n_1 $ and $ n_2 $ vertices each such that there are no adjacent vertices within the same set, being $ n = n_1+n_2 $ the total number of vertices in the bipartite graph. The vertices of the two sets are connected randomly with probability $ p \in [0, 1] $. Another work in this spirit is [18].
We stress that the computational study of the variable symmetric division deg index we perform below is justified by the random nature of the graph models we want to explore. Since a given parameter set $ (n, p) $ [$ (n_1, n_2, p) $] represents an infinite-size ensemble of ER graphs [BR graphs], the computation of $ SDD_\alpha(G) $ on a single graph is irrelevant. In contrast, the computation of $ \left < SDD_\alpha(G) \right > $ (where $ \left < \cdot \right > $ indicates ensemble average) over a large number of random graphs, all characterized by the same parameter set $ (n, p) $ [$ (n_1, n_2, p) $], may provide useful average information about the full ensemble. This computational approach, well known in random matrix theory studies, is not widespread in studies involving topological indices, mainly because topological indices are not commonly applied to random graphs; for very recent exceptions see [19,20,21,22].
From the definition of the variable symmetric division deg index, see Eq (1.1), we have that:
(i) For $ \alpha = 0 $, $ \left < SDD_0(G)\right > $ gives twice the average number of edges of the ER graph. That is,
$ ⟨SDD0(G)⟩=⟨∑uv∈E(G)(d0ud0v+d0vd0u)⟩=⟨∑uv∈E(G)(1+1)⟩=⟨2|E(G)|⟩=n(n−1)p. $
|
(3.1) |
(ii) When $ np\gg 1 $, we can approximate $ d_u \approx d_v \approx \left < d \right > $, then
$ ⟨SDDα(G)⟩≈⟨∑uv∈E(G)(1α+1α)⟩=⟨∑uv∈E(G)2⟩=⟨2|E(G)|⟩=n(n−1)p. $
|
(3.2) |
(iii) By recognizing that the average degree of the ER graph model reads as
$ ⟨d⟩=(n−1)p, $
|
(3.3) |
we can rewrite Eq (3.2) as
$ ⟨SDDα(G)⟩n≈⟨d⟩. $
|
(3.4) |
We stress that Eq (3.4) is expected to be valid for $ np \gg 1 $.
In Figure 2(a) we plot $ \left < SDD_\alpha(G) \right > $ as a function of the probability $ p $ of ER graphs of size $ n = 500 $. All averages in Figure 2 are computed over ensembles of $ 10^7/n $ random graphs. In Figure 2(a) we show curves for $ \alpha\in[0, 4] $. The dashed-magenta curve corresponds to the case $ \alpha = 0 $, which coincides with Eq (3.1). Moreover, we observe that
$ \left < SDD_{\alpha\le 0.5}(G) \right > \approx \left < SDD_0(G) \right > = n(n-1) p \, . $ |
However, once $ \alpha > 0.5 $, the curves $ \left < SDD_\alpha(G) \right > $ versus $ p $ deviate from Eq (3.1), at intermediate values of $ p $, in the form of a bump which is enhanced the larger the value of $ \alpha $ is. Also, in Figure 2(a) we can clearly see that Eq (3.2) is satisfied when $ np\gg 1 $, as expected.
Now, in Figure 2(b) we present $ \left < SDD_\alpha(G) \right > $ as a function of the probability $ p $ of ER graphs of three different sizes. It is clear from this figure that the blocks of curves, characterized by the different graph sizes (and shown in different colors), display similar curves but displaced on both axes. Moreover, the fact that these blocks of curves, plotted in semi-log scale, are shifted the same amount on both $ x- $ and $ y- $axis when doubling $ n $ make us anticipate the scaling of $ \left < SDD_\alpha(G) \right > $. We stress that other average variable degree-based indices on ER random graphs (normalized to the graph size) have been shown to scale with the average degree [22]. Indeed, this statement is encoded in Eq (3.4), that we derived for $ np\gg 1 $ but should serve as the global scaling equation for $ \left < SDD_\alpha(G) \right > $.
Therefore, in Figure 2(c) we show $ \left < SDD_\alpha(G) \right > /n $ as a function of the average degree $ \left < d \right > $ where the same curves of Figure 2(b) have been used. There we verify the global scaling of $ \left < SDD_\alpha(G) \right > $, as anticipated in Eq (3.4), by noticing that the blocks of curves (painted in different colors) for different graph sizes fall on top of each other.
Also, from Figure 2(a)–(c) we observe that the inequality of Theorem 1 is extended to the average variable symmetric division deg index on random graphs:
$ ⟨SDDα(G)⟩≤⟨SDDβ(G)⟩,0<α<β; $
|
(3.5) |
see e.g., the blue arrow in Figure 2(a) which indicates increasing $ \alpha $. Here, the equality is attained if and only if $ p = 1 $. However, we have observed that $ \left < SDD_{ \alpha}(G) \right > \approx \left < SDD_{ \beta}(G) \right > $ already for $ \left < d \right > \ge 10 $.
In Figure 2(d), (e) we present curves of the $ \left < SDD_\alpha(G) \right > $ as a function of the probability $ p $ of BR graphs. For simplicity we show results for BR graphs composed by sets of equal sizes $ n_1 = n_2 $. In Figure 2(d) we consider the case of $ n_1 = n_2 = 500 $ while in (e) we report $ n_1 = n_2 = \{125, 250, 500\} $. In both figures we show curves for $ \alpha\in[0, 4] $ in steps of $ 0.5 $.
Since edges in a bipartite graph join vertices of different sets, and we are labeling here the sets as sets 1 and 2, we replace $ d_u $ by $ d_1 $ and $ d_v $ by $ d_2 $ in the expression for the $ SDD_\alpha(G) $ index below. Thus,
(i) For $ \alpha = 0 $, $ \left < SDD_0(G)\right > $ gives twice the average number of edges of the BG graph. That is,
$ ⟨SDD0(G)⟩=⟨∑E(G)(d01d02+d02d01)⟩=⟨∑E(G)(1+1)⟩=⟨2|E(G)|⟩=2n1n2p. $
|
(3.6) |
(ii) When both $ n_1p\gg 1 $ and $ n_2p\gg 1 $, we can approximate $ d_1 \approx \left < d_1 \right > $ and $ d_2 \approx \left < d_2 \right > $, then
$ ⟨SDDα(G)⟩≈⟨∑E(G)(⟨d1⟩α⟨d2⟩α+⟨d2⟩α⟨d1⟩α)⟩=⟨|E(G)|(⟨d1⟩α⟨d2⟩α+⟨d2⟩α⟨d1⟩α)⟩. $
|
(3.7) |
(iii) In the case we consider in Figure 2(d)–(f), where $ n_1 = n_2 = n/2 $, so that $ \left < d_1 \right > = \left < d_2 \right > = \left < d \right > $, Equation (3.7) reduces to
$ ⟨SDDα(G)⟩≈⟨2|E(G)|⟩=2n1n2p=n22p. $
|
(3.8) |
(iv) By recognizing that $ \left < d \right > = np/2 $ we can rewrite Eq (3.8) as
$ ⟨SDDα(G)⟩n≈⟨d⟩. $
|
(3.9) |
We stress that Eq (3.9) is expected to be valid for $ np \gg 1 $. We also note that Eq (3.9) has exactly the same form as Eq (3.4).
From Figure 2(d), (e) we note that
$ \left < SDD_{\alpha\le 0.5}(G) \right > \approx \left < SDD_0(G) \right > = 2n_1n_2 p \, , $ |
see the dashed-magenta curve in Figure 2(d). But once $ \alpha > 0.5 $, the curves $ \left < SDD_\alpha(G) \right > $ versus $ p $ deviate from Eq (3.6), at intermediate values of $ p $, in the form of bumps which are enhanced the larger the value of $ \alpha $ is. These bumps make clear the validity of inequality (3.5) on BR graphs; see e.g., the blue arrow in Figure 2(d) which indicates increasing $ \alpha $.
Finally, following the scaling analysis made in the previous subsection for ER graphs, in Figure 2(f) we plot the $ \left < SDD_\alpha(G) \right > /n $ as a function of the average degree $ \left < d \right > $ where the same data sets of Figure 2(e) have been used. Thus we verify that $ \left < SDD_\alpha(G) \right > $/n scales with $ \left < d \right > $, as anticipated in Eq (3.9); that is, the blocks of curves (painted in different colors) for different graph sizes coincide.
In this paper we have performed analytical and computational studies of the variable symmetric division deg index $ SDD_\alpha(G) $. First, we provided a monotonicity property and obtained new inequalities connecting $ SDD_\alpha(G) $ with other well–known topological indices such as the variable inverse sum deg index, as well as the the modified Narumi-Katayama index. Then, we apply the index $ SDD_\alpha(G) $ on two ensembles of random graphs: Erd\H{o}s-Rényi graphs and bipartite random graphs. Thus, we computationally showed, for both random graph models, that the ratio $ \langle SDD_\alpha(G) \rangle/n $ is a function of the average degree $ \langle d \rangle $ only ($ n $ being the order of the graph). We note that this last result, also observed for other variable topological indices [22], is valid for random bipartite graphs only when they are formed by sets of the same size.
Since many important topological indices can be written as
$ X(G) = \sum\limits_{uv \in E(G)} F(d_u, d_v), $ |
an open problem is to extend the results of this paper to other indices of this kind.
J. A. M.-B. has been supported by CONACyT (Grant No. CB-2016-286633-F). J. M. R. and J. M. S. have been supported by a grant from Agencia Estatal de Investigación (PID2019-106433GBI00/AEI/10.13039/501100011033), Spain. J. M. R. has been supported by the Madrid Government (Comunidad de Madrid-Spain) under the Multiannual Agreement with UC3M in the line of Excellence of University Professors (EPUC3M23), and in the context of the V PRICIT (Regional Programme of Research and Technological Innovation).
The authors declare there is no conflict of interest.
[1] | Buzsaki G, Draguhn A. (2004) Neuronal oscillations in cortical networks. Science 304(5679):1926-9. |
[2] | Albensi BC. (2007) The NMDA receptor/ion channel complex: a drug target for modulating synaptic plasticity and excitotoxicity. Curr Pharm Des 13(31): 3185-94. |
[3] | Seeburg PH. (1994) Molecular biology of NMDA receptors. In: Watkins JC, Collingridge GL. Authors, The NMDA receptor, 2 Eds. , New York: Oxford University Press. |
[4] | Gibb AJ. (1994) Activation of NMDA receptors. In: Watkins JC, Collingridge GL. Authors, The NMDA receptor, 2 Eds. , New York: Oxford University Press. |
[5] | Alford S, Brodin L. (1994) The role of NMDA receptors in synaptic integration and the organization of motor patterns. In: Watkins JC, Collingridge GL. Authors, The NMDA receptor, 2 Eds. , New York: Oxford University Press. |
[6] | Morris RGM, Davis M. (1994) The role of NMDA receptors in learning and memory. In: Watkins JC, Collingridge GL. Authors, The NMDA receptor, 2 Eds. , New York: Oxford University Press. |
[7] | Garthwaite J. (1994) NMDA receptors, neuronal development, and neurodegeneration. In: Watkins JC, Collingridge GL. Authors, The NMDA receptor, 2 Eds. , New York: Oxford University Press. |
[8] | Bear MF, Malenka RC. (1994) Synaptic plasticity: LTP and LTD. Curr Opin Neurobiol 4(3):389-99. |
[9] | Möddel G, Jacobson B, Ying Z, et al. (2005) The NMDA receptor NR2B subunit contributes to epileptogenesis in human cortical dysplasia. Brain Res 1046(1-2): 10-23. |
[10] | Mody I, MacDonald JF. (1995) NMDA receptor-dependent excitotoxicity: the role of intracellular Ca2+ release. Trends Pharmacol Sci 16(10): 356-9. |
[11] | Nagle R, Saff B, Snider A. (2012) Fundamentals of differential equations and boundary value problems. 6 Eds. , Boston: Addison-Wesley. |
[12] | Izhikevich EM. (2007) Dynamical systems in neuroscience: the geometry of excitability and bursting. Computational neuroscience, Cambridge: MIT Press, 441. |
[13] | Koch C, Segev I. (1998) Methods in neuronal modeling: From ions to networks, Cambridge, Massachusetts: MIT Press. |
[14] | Kuramoto Y. (1984) Chemical oscillations, waves, and turbulence. In: Springer series in synergetics of Berlin, New York: Springer-Verlag, 156. |
[15] | Niedermeyer E, Schomer DL, Lopes da Silva FH. (2011) Niedermeyer's electroencephalography: basic principles, clinical applications, and related fields. 6 Eds. , Philadelphia: Wolters Kluwer Health/Lippincott Williams & Wilkins, 1275. |
[16] | Tass PA. (2007) Phase resetting in medicine and biology stochastic modelling and data analysis, In: Springer Series in Synergetics, Berlin: Springer-Verlag. |
[17] | Albensi BC, et al. (2004) Why do many NMDA antagonists fail, while others are safe and effective at blocking excitotoxicity associated with dementia and acute injury? Am J Alzheimers Dis Other Demen 19(5): 269-74. |
[18] | Bliss TV, Collingridge GL. (1993) A synaptic model of memory: long-term potentiation in the hippocampus. Nature 361(6407): 31-9. |
[19] | Lynch MA. (2004) Long-term potentiation and memory. Physiol Rev 84(1): 87-136. |
[20] | Lisman J. (2003) Long-term potentiation: outstanding questions and attempted synthesis. Philos Trans R Soc Lond B Biol Sci 358(1432): 829-42. |
[21] | Malenka RC. (1994) Synaptic plasticity in the hippocampus: LTP and LTD. Cell 78(4): 535-8. |
[22] | Yang SN, Tang YG, Zucker RS. (1999) Selective induction of LTP and LTD by postsynaptic [Ca2+]i elevation. J Neurophysiol 81(2): 781-7. |
[23] | Huang YY, Malenka RC. (1993) Examination of TEA-induced synaptic enhancement in area CA1 of the hippocampus: the role of voltage-dependent Ca2+ channels in the induction of LTP. J Neurosci 13(2): 568-76. |
[24] | Abraham WC, Williams JM. (2003) Properties and mechanisms of LTP maintenance. Neuroscientist 9(6): 463-74. |
[25] | Chen HS, Lipton SA. (2006) The chemical biology of clinically tolerated NMDA receptor antagonists. J Neurochem 97(6): 1611-26. |
[26] | McIlhinney RA, et al. (2003) Assembly of N-methyl-D-aspartate (NMDA) receptors. Biochem Soc Trans 31(Pt 4): 865-8. |
[27] | Waxman EA, Lynch DR. (2005) N-methyl-D-aspartate receptor subtypes: multiple roles in excitotoxicity and neurological disease. Neuroscientist 1(1): 37-49. |
[28] | Dingledine R, et al. (1999) The glutamate receptor ion channels. Pharmacol Rev 51(1): 7-61. |
[29] | Lynch DR, Guttmann RP. (2001) NMDA receptor pharmacology: perspectives from molecular biology. Curr Drug Targets 2(3): 215-31. |
[30] | Farber NB, Newcomer JW, Olney JW. (1998) The glutamate synapse in neuropsychiatric disorders, progress in brain research. In: Ottersen OP, Langmoen IA, Gjerstad L. Authors, The gluatmate synapse as a therapeutic target: molecular organization and pathology of the glutamate synapse, New York: Elsevier. |
[31] | Lindsley CW, et al. (2006) Progress towards validating the NMDA receptor hypofunction hypothesis of schizophrenia. Curr Top Med Chem 6(8): 771-85. |
[32] | Ikonomidou C, Turski L. (2002) Why did NMDA receptor antagonists fail clinical trials for stroke and traumatic brain injury? Lancet Neurol 1(6): 383-6. |
[33] | Hardingham GE, Bading H. (2003) The Yin and Yang of NMDA receptor signalling. Trends Neurosci 26(2): 81-9. |
[34] | Whittington MA, Traub RD, Jefferys JG. (1995) Synchronized oscillations in interneuron networks driven by metabotropic glutamate receptor activation. Nature 373(6515): 612-5. |
[35] | Middleton S, et al. (2008) NMDA receptor-dependent switching between different gamma rhythm-generating microcircuits in entorhinal cortex. Proc Natl Acad Sci USA 105(47): 18572-7. |
[36] | Lazarewicz MT, et al. (2010) Ketamine modulates theta and gamma oscillations. J Cogn Neurosci 22(7): 1452-64. |
[37] | Cabral HO, et al. (2014) Oscillatory dynamics and place field maps reflect hippocampal ensemble processing of sequence and place memory under NMDA receptor control. Neuron81(2): 402-15. |
[38] | Jacobsen RB, Ulrich D, Huguenard JR. (2001) GABA(B) and NMDA receptors contribute to spindle-like oscillations in rat thalamus in vitro. J Neurophysiol 86(3): 1365-75. |
[39] | Korotkova T, et al. (2010) NMDA receptor ablation on parvalbumin-positive interneurons impairs hippocampal synchrony, spatial representations, and working memory. Neuron 68(3):557-69. |
[40] | Carlen M, et al. (2012) A critical role for NMDA receptors in parvalbumin interneurons for gamma rhythm induction and behavior. Mol Psychiatry 17(5): 537-48. |
[41] | Anver H, et al. (2010) NMDA receptor hypofunction phase couples independent gamma-oscillations in the rat visual cortex. Neuropsychopharmacology 36(2): 519-28. |
[42] | van Wingerden M, et al. (2012) NMDA receptors control cue-outcome selectivity and plasticity of orbitofrontal firing patterns during associative stimulus-reward learning. Neuron 76(4):813-25. |
[43] | Wallen P, Grillner S. (1987) N-methyl-D-aspartate receptor-induced, inherent oscillatory activity in neurons active during fictive locomotion in the lamprey. J Neurosci 7(9): 2745-55. |
[44] | Wang D, Grillner S, Wallen P. (2013) Calcium dynamics during NMDA-induced membrane potential oscillations in lamprey spinal neurons, contribution of L-type calcium channels (CaV1. 3). J Physiol 591(Pt 10): 2509-21. |
[45] | Lee S, Sen K, Kopell N. (2009), Cortical gamma rhythms modulate NMDAR-mediated spike timing dependent plasticity in a biophysical model. PLoS Comput Biol 5(12): e1000602. |
[46] | McNaughton BL, et al. (2006) Path integration and the neural basis of the "cognitive map". Nat Rev Neurosci 7(8): 663-78. |
[47] | Burgess N. (2006) Spatial memory: how egocentric and allocentric combine. Trends Cogn Sci10(12): 551-7. |
[48] | O'Keefe J, Nadel L. (1978) The Hippocampus as a Cogntive Map. London: University of Oxford Press. |
1. | Bahati Erick, Maranya Mayengo, Modelling the dynamics of Cassava Mosaic Disease with non-cassava host plants, 2022, 33, 23529148, 101086, 10.1016/j.imu.2022.101086 | |
2. | Caihong Song, Ning Li, Dynamic analysis and bifurcation control of a fractional-order cassava mosaic disease model, 2022, 1598-5865, 10.1007/s12190-022-01809-9 | |
3. | Nabeela Anwar, Shafaq Naz, Muhammad Shoaib, Reliable numerical treatment with Adams and BDF methods for plant virus propagation model by vector with impact of time lag and density, 2022, 8, 2297-4687, 10.3389/fams.2022.1001392 | |
4. | Geofrey Sikazwe, Rosita E.E. Yocgo, Pietro Landi, David M. Richardson, Cang Hui, Managing whitefly development to control cassava brown streak virus coinfections, 2024, 493, 03043800, 110753, 10.1016/j.ecolmodel.2024.110753 | |
5. | Tariq Q. S. Abdullah, Gang Huang, Wadhah Al-Sadi, Yasser Aboelmagd, Wael Mobarak, Fractional Dynamics of Cassava Mosaic Disease Model with Recovery Rate Using New Proposed Numerical Scheme, 2024, 12, 2227-7390, 2386, 10.3390/math12152386 | |
6. | Sireepatch Sangsawang, Usa Humphries, Amir Khan, Puntani Pongsumpun, Sensitivity analysis and optimal control for the dynamic mathematical model of cassava mosaic disease, 2024, 14, 2158-3226, 10.1063/5.0207497 | |
7. | E. Azroul, N. Kamali, M. Shimi, Novel insights into Cassava mosaic disease using Caputo fractional derivative: modeling and analysis, 2025, 0003-6811, 1, 10.1080/00036811.2025.2509314 |