This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Mathematical representation of the structure of neuron-glia networks

Marco Peña-Garcia, Francesco Peña-Garcia, Walter Cabrera-Febola
Group of Natural Structures and Theoretical Research
Universidad Nacional Mayor de San Marcos
Lima, Perú
[email protected]
&Nelson Castro
Department of Model Risk Management
Banco de Crédito del Perú
Lima, Perú
Abstract

Network representations of the nervous system have been useful for the understanding of brain phenomena such as perception, motor coordination, and memory. Although brains are composed of both neurons and glial cells, neuron-glial networks have been little studied so far. Given the emergent role of glial cells in information transmission in the brain, we developed a mathematical representation for neuron-glial networks (Υ\Upsilon-graph). We also defined the concepts of isomorphisms, unnested form (multidigraph) and matrix equation for Υ\Upsilon-graphs. Although we found several network motives where the isomorphism between unnested forms does not guarantees the isomorphism between their respective Υ\Upsilon-graphs, we found that if the matrix equations satisfy some conditions, the unnested forms isomorphism guarantees the isomorphism between Υ\Upsilon-graphs. Finally, we introduced a novel approach to modeling the network shape. Our work presents a mathematical framework for working with neuron-glia networks.

Keywords Neuron-glia network, Multiset, Graph theory

1 Introduction

Traditionally, it was commonly accepted that only neurons are involved in the brain information processing while glial cells are relegated to support and protect them. A more updated view suggests that glial cells are also involved in the information processing. In fact, one type of glia cell - astrocytes - couples to single synapses between neurons, forming tripartite synapses [15]. Astrocytes have an active role at tripartite synapses. On one hand, they can directly readout signals from the presynaptic neuron (e.g. they express receptors for GABA, dopamine, glutamate, and glycine)[18, 36]. On the other hand, they can modulate the activity of the postsynaptic neuron via the release of neuroactive substances called gliotransmitters [15, 29].

Neuron-driven activation of astrocytes induces an increase in cytosolic calcium [18, 37]. This increase in calcium, in turn, triggers the release of gliotransmitters [8]. More importantly, local increases in calcium spreads through the astrocyte syncytium, i.e. astrocyte network, mainly via IP3 diffusion [11, 30]. Interestingly, recent studies suggest that the astrocyte network plays a role in neural network phenomena such as -burst synchronization, bursting behavior, barrage firing, and gamma waves- and in cognitive capabilities (e.g. recognition memory) [6, 10, 21]. Therefore, studying neuron-glia networks could be critical to understand the brain dynamics of large populations of neurons.

Although neuron-glia networks have been modelled before, finding that astrocytes modulate dynamic coordination, facilitate ultra-slow oscillation, and increase the occurrence of bursting-like spikes, they lack a formal network representation [5, 34, 39]. Neural networks are typically represented using graphs or digraphs [2, 26, 32]. These representations allow the study of features such as connectivity, modularity, and centrality. [2]. So far, graph-like structures has guided the study of connectomes at all scales [2, 20, 33]. However, neuron-glia networks have tripartite synapses that relate three cells at once, thus they can not be represented by edges of a graph nor a digraph.

In this paper, we addressed this problem by defining a representations of neuron-glia networks. In order to achieve this goal, first, we defined a synaptic configuration as a set of ordered pairs that represent the information flux between cells in a synapse. Then, we represented a collection of synapses as a Υ\Upsilon-family that relates synaptic configurations to indexes (single synapses), in order to distinguish synapses that share a synaptic configuration. To simplify this representation, we defined Υ\Upsilon-multisets that contain identical copies of repeated synaptic configuration. A set of neuron-like cells (neurons and astrocytes) together with an Υ\Upsilon-multiset form an Υ\Upsilon-graph which represent a neuron-glia network. This representation allowed us to adapt the concept of multidigraph isomorphism to neuron-glia networks. Then, we defined the ϕ\phi function that maps Υ\Upsilon-graphs with their unnested forms (multidigraphs) which contain only ordered pairs. Using this function, we found network motifs (propositions 2-5) where the isomorphism between unnested forms does not guarantee the isomorphism between Υ\Upsilon-graphs, meaning misrepresentation by the multidigraphs. However, we also found a mathematical condition (proposition 6) that allows the unnested forms isomorphism to guarantee the Υ\Upsilon-graphs isomorphisms. Finally, we introduced a novel approach to model neuron-glia network shape, using Υ\Upsilon-families, natural structure theory, and topology theory [3, 4]. Future research could use our definition of Υ\Upsilon-graph isomorphism as part of an algorithm that find neuron-glia network motifs in a connectome that includes astrocytes such as the connectome reconstructed by Sizis et al., 2021 [42] (see [23] for examples of motif analysis of a connectome). Another research avenue would be extending graph measures (e.g. centrality, modularity, and distance) to Υ\Upsilon-graphs, in order to analyze connectomics data (see [23] for examples of use of network metrics to study a connectome) or to study how these measures correlates with different predicted networks dynamics in dynamical models.

2 Basic definitions

We follow the definitions and notations of [13] and [14] about multisets (msets) and related objects. We also follow the concepts of [1] about digraphs and related objects. In the following formal definitions, we summarize the concepts about multisets and graph theory that are necessary to understand our theoretical framework.

Definition 1.

[14] A collection of elements containing duplicates is called a multiset. Formally, if XX is a set of elements, a multiset MM drawn from the set XX is represented by a function CountMCount_{M} defined as CountM:X0Count_{M}:X\to\mathbb{N}_{0} where 0\mathbb{N}_{0} represents the set of non negative integers.

CountMCount_{M} associates each element of the set XX with its occurrence in the multiset MM [14, 13]. The support set of MM is denoted by MM^{*} and it is defined as the subset of XX in which the occurrence of all its element in MM is greater than 0 [13].

Definition 2.

Let two digraphs D=(V(D),A(D))D=(V(D),A(D)) and H=(V(D),A(D))H=(V(D),A(D)) where V()V() and A()A() denote the set of vertices and arcs of the digraph, respectively. A digraph homomorphism is a function f:V(D)V(H)f:V(D)\to V(H) if it preserves arcs, meaning that (x,y)A(D)(x,y)\in A(D) implies (f(x),f(y))A(H)(f(x),f(y))\in A(H).

The following definition formalizes the concept of directed pseudograph [1] to create our multidigraph definition, using multiset theory [14, 13].

Definition 3.

A finite multidigraph QQ is a tuple (V(Q),A(Q))(V(Q),A(Q)) where V(Q)V(Q) is a finite vertex set and A(Q)A(Q) is the finite arcs multiset with support set A(Q)V(Q)×V(Q)A(Q)^{*}\subset V(Q)\times V(Q).

Note that if all edges appear once, A(Q)A(Q) is equivalent to a set, and the multidigraph QQ is equivalent to a digraph (loop allowed). Note that A(Q)A(Q)^{*} is a binary relation on V(Q)V(Q). We also formalize the multidigraphs homomorphism in the context of multiset theory as follows.

Definition 4.

Given two multidigraphs QQ and PP with vertex sets V(Q)V(Q) and V(P)V(P), respectively, an homomorphism is a bijective function f:V(Q)V(P)f:V(Q)\to V(P) such that for any (u,v)A(Q)(u,v)\in A(Q)^{*} the following is true.

CountA(Q)((u,v))=CountA(P)((f(u),f(v)))\text{Count}_{A(Q)}((u,v))=\text{Count}_{A(P)}((f(u),f(v)))

If QQ is homomorphic to PP, and PP is homomorphic to QQ, QQ and PP are isomorphic (QPQ\cong P).

3 Theoretical framework

Like neurons, astrocytes receive and release neurotransmitters, so we will refer to both as neuron-like cells. In the following subsection, NN and AA will represent a set of neuron and a set of astrocytes, respectively. Both NN and AA are sets of neuron-like cells (NL-sets). Additionally, we can define a NL-set XX as the union of NN and AA (X=NAX=N\cup A). Since astrocyte and neurons are different cellular types [15], a set of neurons and a set of astrocytes are always disjoint sets (i.e., NA=N\cap A=\varnothing).

3.1 Synaptic configurations

Synapses allow the information flux from one neuron-like cell to another. In particular, tripartite synapses allow a directed information flux from the pre-synaptic neuron to the post-synaptic neuron, from the pre-synaptic neuron to an astrocyte, and from the astrocyte to the post-synaptic neuron. We can represent this synaptic configuration as a set of ordered pairs (binary relation) where each ordered pair represents a direction of information flux.

Definition 5.

A tripartite configuration is represented by sts_{t} that is set containing three ordered pairs. The ordered pairs represent the information flow of a tripartite synapse. If nin_{i} is the presynaptic neuron, njn_{j} is the postsynaptic neuron, and aia_{i} is an astrocyte, sts_{t} is defined as

st:={(ni,nj),(ni,ai),(ai,nj)}.s_{t}:=\{(n_{i},n_{j}),(n_{i},a_{i}),(a_{i},n_{j})\}.

Since not all directed synapses between neurons are necessarily associated with astrocytes, we define a type of configuration called one-one-directed configuration in the following.

Definition 6.

A one-one-directed configuration is represented by sds_{d} that is a set containing only one ordered pair. If nin_{i} is the presynaptic neuron and njn_{j} is the postsynaptic neuron, sds_{d} is defined as

sd:={(ni,nj)}.s_{d}:=\{(n_{i},n_{j})\}.

Gap junctions can also connect neurons and astrocytes [11, 24, 30]. Gap junctions mainly allow the diffusion of ions and small molecules between neurons or between astrocytes [11, 24, 30]. Since gap junctions bidirectionally connect neural-like cells, we represent their synaptic configurations as symmetric binary relations. We will call this synaptic configuration symmetric configuration instead of gap junction configuration to highlight that the information flow is bidirectional.

Definition 7.

A symmetric configuration is represented by sss_{s} that is set containing two ordered pairs. If two neuron-like cells, xix_{i} and xjx_{j}, from the same cellular type (i.e. xi,xjN or xi,xjAx_{i},x_{j}\in N\textrm{ or }x_{i},x_{j}\in A) are connected by a gap junction, sss_{s} is defined as

ss:={(xi,xj),(xj,xi)}.s_{s}:=\{(x_{i},x_{j}),(x_{j},x_{i})\}.

All these synaptic configuration are graphically explained in fig 1.

Note 1.

In an ordered pair (a,b)(a,b), aa is the first coordinate and bb is the second coordinate.

3.2 Υ\Upsilon-graph representation

Here, we define two representations of the collection of synapses of a neuron-glia network: Υ\Upsilon-family and Υ\Upsilon-multiset. But first, we define a set of synaptic configurations SXS_{X} on the NL-set XX of a neuron-glia network. For the sake of clarity, the subscript of a set of synaptic configurations refers to the NL-set on which the synaptic configurations are defined.

Definition 8.

A set of synaptic configurations SXS_{X} is a finite collection of sets of ordered pairs within X×XX\times X, where XX is a NL-set, satisfying the following conditions:

A1. All elements of SXS_{X} are synaptic configurations sds_{d} (one-one-directed configuration), sss_{s} (symmetric configuration), or sts_{t} (tripartite configuration).

A2. There is at least one tripartite configuration sts_{t} in SxS_{x}.

A3. All neuron-like cells of XX are a coordinate of an ordered pair that belong to at least one synaptic configuration of SXS_{X}.

Note 2.

A2 defines the minimal composition of a NL-set in the context of neuron-glia networks. This condition is necessary because there must be, at least, one tripartite synapses in a neuron-glia network. A3 guarantees that the elements of XX participate in at least one synaptic configuration of SXS_{X}. In other words, all neuron-like cells form at least one synapse.

The ordered pairs within the elements of SXS_{X} represent the directions of information flow inside a single synapse (see section 5 for a biological definition of single synapse or see fig 1a for a diagram); however, in a network, single synapses can share a synaptic configuration. To represent this fact, we define the following concept.

Definition 9.

An Υ\Upsilon-family FF is a finite indexed family F:ISXF:I\to S_{X} where II is the set of single synapses of XX (II\neq\varnothing) such that F(a)=bF(a)=b if and only if bb is the synaptic configuration of aa.

Υ\Upsilon-families relate single synapses with their synaptic configurations. Consequently, two single synapses ii and jj with synaptic configuration ss can be represented by the equations F(i)=sF(i)=s and F(j)=sF(j)=s. In other words, Υ\Upsilon-families are functions from real connections (single synapses) to mathematical objects (synaptic configurations). To facilitate the comparison of networks, we define the concept of Υ\Upsilon-multiset which is a multiset represented by a count function that count how many single synapses share a synaptic configuration in an Υ\Upsilon-family.

Definition 10.

An Υ\Upsilon-multiset AxA_{x} is a mset represented by a function Count:SX+\text{Count}:S_{X}\to\mathbb{Z}^{+}, satisfying the following condition : USX\forall U\in S_{X}, CountAx(U)={i:F(i)=U}Count_{A_{x}}(U)=\textit{}\mid\!\{i:F(i)=U\}\!\mid where FF is an Υ\Upsilon-family ISXI\to S_{X}.

Since {i:F(i)=U}\{i:F(i)=U\} is the set of preimages under FF of UU, its cardinality represents the number of occurrence of a synaptic configurations in the network. Now, given definition 10, we can define an Υ\Upsilon-graph that represent a neuron-glia network as follows.

Definition 11.

An Υ\Upsilon-graph is a tuple (X,Ax)(X,A_{x}) where XX is a NL-set and AxA_{x} is an Υ\Upsilon-multiset represented by a function Count:SX+\text{Count}:S_{X}\to\mathbb{Z}^{+}.

A Υ\Upsilon-graph (X,Ax)(X,A_{x}) is called connected Υ\Upsilon-graph if for every different pair of elements x,yXx,y\in X, there exists sequence (an)0k(a_{n})_{0}^{k} where each pair of consecutive elements is part of a some relation (set of ordered pairs) rAxr\in A_{x} and a0=xak=ya_{0}=x\land a_{k}=y.

Refer to caption
Figure 1: a) Schematic representation of a neuron-glia network. N1N_{1} and N2N_{2} are two neurons, and A1A_{1} and A2A_{2} are two astrocytes. 1 is a tripartite synapse, 2 is a gap junction synapse between neurons, 3 is an one-one-directed synapse, and 4 is a gap junction between astrocytes. We show an additional non-labeled tripartite synapse to highlight that two single synapses can share the same synaptic configuration. b) Graphical representation of the directions of information flux inside synapses (synaptic configuration) (numbers refer to single synapses showed in a))

3.3 Υ\Upsilon-graph isomorphism

Similarly to multidigraph homomorphism (definition 4), we define homomorphism for Υ\Upsilon-graphs in definition 12. But, first, for the sake of clarity, we introduce the following notation, being ff a bijective function between two NL-sets and ss a set of ordered pairs of elements of the domain of ff, if s={(a,b)}s=\{(a,b)\}, s={(a,b),(a,b)}s=\{(a,b),(a,b)\}, or s={(a,b),(b,c),(a,c)}s=\{(a,b),(b,c),(a,c)\}, then, sfs_{f} is {(f(a),f(b))}\{(f(a),f(b))\}, {(f(a),f(b)),(f(b),f(a))}\{(f(a),f(b)),(f(b),f(a))\}, or {(f(a),f(b)),\{(f(a),f(b)), (f(b),f(c)),(f(b),f(c)), (f(a),f(c))}(f(a),f(c))\}, in that order. It is clear that ss and sfs_{f} can be associated with an injective function; however, such a function will be omitted for simplicity.

Definition 12.

Given two Υ\Upsilon-graphs α=(X,Ax)\alpha=(X,A_{x}) and β=(Y,Ay)\beta=(Y,A_{y}), the bijective function f:XYf:X\to Y is an homomorphism if, for any element ss of AxA_{x}^{*}, the following condition is true.

CountAx(s)=CountAy(sf)\text{Count}_{A_{x}}(s)=\text{Count}_{A_{y}}(s_{f})

As before, if α\alpha is homomorphic to β\beta and β\beta is homomorphic to α\alpha, α\alpha and β\beta are isomorphic (αβ\alpha\cong\beta).

In order to add functional relevance to the isomorphism concept, we create the term α\alpha-isomorphic (represented as α\cong_{\alpha}) which is defined as follows.

Definition 13.

The Υ\Upsilon-graphs α=(X,Ax)\alpha=(X,A_{x}) and β=(Y,Ay)\beta=(Y,A_{y}) are α\alpha-isomorphic if the homomorphisms from α\alpha to β\beta and from β\beta to α\alpha only relate neurons to neurons and astrocytes to astrocytes.

This concept will also be applied to multidigraph isomorphisms if the vertex sets contains neurons and astrocytes. So, if xαyx\cong_{\alpha}y, then the homomorphisms from xx to yy and from yy to xx are bijective functions that relate neurons to neurons and astrocytes to astrocytes.

3.4 Unnesting an Υ\Upsilon-graph

Every Υ\Upsilon-multiset is based on a NL-set (definition 10) and, for every Υ\Upsilon-multiset, there is only one associated NL-set (see appendix 9.1, proposition 7). Clearly, this implies that for a given Υ\Upsilon-multiset AxA_{x}, there exists one and only one associated connected Υ\Upsilon-graph (X,Ax)(X,A_{x}). Now, let preϕpre-\phi be a function that relates Υ\Upsilon-multisets with multisets of all the ordered pairs of all elements of the Υ\Upsilon-multiset (see appendix 9.2 for a detailed definition). Similarly to proposition proposition 7, we show that there exists only one NL-set per preϕ(Ax)pre-\phi(A_{x}) for any AxA_{x} in proposition 8 (see appendix 9.3). Therefore, for a given multiset preϕ(Ax)pre-\phi(A_{x}), there exists one and only one associated multidigraph (X,preϕ(Ax))(X,pre-\phi(A_{x})). Using preϕpre-\phi, it is possible to construct another function called ϕ\phi that relates Υ\Upsilon-graphs to multidigraphs as it is shown in the following definition.

Definition 14.

Being 𝐏\mathbf{P} a non-empty set of Υ\Upsilon-graphs and 𝐐\mathbf{Q} a large enough non-empty set of multidigraphs, ϕ\phi is a function defined as

ϕ:𝐏𝐐xy=ϕ(x)\begin{array}[]{cccc}\phi:&\mathbf{P}&\to&\mathbf{Q}\\ &x&\mapsto&y=\phi(x)\end{array} (1)

if x:=(X,Ax)x:=(X,A_{x}) and y:=(V(y),E(y))y:=(V(y),E(y)), then yy is the image of xx, i.e., y=ϕ(x)y=\phi(x), if and only if X=V(y)X=V(y) and E(y)=preϕ(Ax)E(y)=pre-\phi(A_{x}).

This new function would represent the “unnested relation”, being the image the unnested form of the argument. Fig. 2 graphically explains the domain and range of ϕ\phi.

Refer to caption
Figure 2: Venn diagram of a set of Υ\Upsilon-graphs (𝐏\mathbf{P}) and a set of their associated multidigraphs (𝐐\mathbf{Q}), as well as the function ϕ\phi that connects them. 𝐏=(x,y,)\mathbf{P}=(x,y,...) and 𝐐=(ϕ(x),ϕ(y),)\mathbf{Q}=(\phi(x),\phi(y),...)

3.5 Connecting Υ\Upsilon-graphs and their unnested forms through matrices

Let x:=(X,Ax)x:=(X,A_{x}) be any Υ\Upsilon-graph and ϕ(x):=(X,preϕ(Ax))\phi(x):=(X,pre-\phi(A_{x})), its unnested form. The following proposition 1 establishes a connection between xx and ϕ(x)\phi(x) in terms of repetition of synaptic configurations and ordered pairs.

Proposition 1.

Let αpreϕ(Ax)\alpha\in pre-\phi(A_{x}) be any vertex of ϕ(x)\phi(x) (i.e., an ordered pair), if s1,s2,,sks_{1},s_{2},...,s_{k} are all synaptic configurations (elements of AxA_{x}) that include α\alpha (αsi\alpha\in s_{i} for all i[1,k]i\in[1,k]\cap\mathbb{Z}), then

Countpreϕ(Ax)(α)=CountAx(s1)+CountAx(s2)++CountAx(sk)Count_{pre-\phi(A_{x})}(\alpha)=Count_{A_{x}}(s_{1})+Count_{A_{x}}(s_{2})+...+Count_{A_{x}}(s_{k}) (2)

Proposition 1 states that, for any ordered pairs of the unnested form (multidigraph), the number of repetitions of the ordered pair is the summation of the number of its occurrences across all synaptic configurations in which it is present. This is formally proven in appendix 9.5.

If we assemble all the equations of the form (2) of all the elements of preϕ(Ax)pre-\phi(A_{x})^{*}, then we form a system of equations that works as a more intuitive bridge between xx and ϕ(x)\phi(x) than ϕ\phi itself. If preϕ(Ax)pre-\phi(A_{x})^{*} has mm elements, the system of equations takes the following form.

{Countpreϕ(Ax)(α1)=CountAx(a1)++CountAx(ak)Countpreϕ(Ax)(α2)=CountAx(b1)++CountAx(bk)Countpreϕ(Ax)(αm)=CountAx(z1)++CountAx(zk′′)\begin{cases}Count_{pre-\phi(A_{x})}(\alpha_{1})=Count_{A_{x}}(a_{1})+...+Count_{A_{x}}(a_{k})\\ Count_{pre-\phi(A_{x})}(\alpha_{2})=Count_{A_{x}}(b_{1})+...+Count_{A_{x}}(b_{k^{{}^{\prime}}})\\ ...\\ Count_{pre-\phi(A_{x})}(\alpha_{m})=Count_{A_{x}}(z_{1})+...+Count_{A_{x}}(z_{k^{{}^{\prime\prime}}})\\ \end{cases} (3)

where α1a1\alpha_{1}\in a_{1}, …, α1ak\alpha_{1}\in a_{k}, α2b1\alpha_{2}\in b_{1}, and so on. Take not that a1a_{1} is not necessarily different from b1b_{1}. We can go a step further to order the system of equations 3 and hence facilitate the operations. In fact, if AxA_{x}^{*} has nn elements (s1,s2,,sns_{1},s_{2},...,s_{n}), then the system of equations is equivalent to

{Countpreϕ(Ax)(α1)=c11CountAx(s1)++c1nCountAx(sn)Countpreϕ(Ax)(αm)=cm1CountAx(s1)++cmnCountAx(sn)\begin{cases}Count_{pre-\phi(A_{x})}(\alpha_{1})=c_{11}Count_{A_{x}}(s_{1})+...+c_{1n}Count_{A_{x}}(s_{n})\\ ...\\ Count_{pre-\phi(A_{x})}(\alpha_{m})=c_{m1}Count_{A_{x}}(s_{1})+...+c_{mn}Count_{A_{x}}(s_{n})\\ \end{cases} (4)

Where each cijc_{ij} represents the presence (cij=1c_{ij}=1) or absence (cij=0c_{ij}=0) of the ii-th ordered pair (αi\alpha_{i}) in the synaptic configuration sjs_{j} . It is possible to represent this later system of equations 4 with the following matrix equation:

P=AxP=Ax (5)

which is amenable for any mathematical treatment. This matrix equation is called the Υ\Upsilon-graph-multidigraph equation or Υ\Upsilon-gm equation.

The vector PP is a mm by 1 matrix. Each element represents the number of occurrences of an ordered pair in preϕ(Ax)pre-\phi(A_{x}) (Countpreϕ(Ax)Count_{pre-\phi(A_{x})}). The coefficient matrix AA is a mm by nn matrix with 1 and 0 as elements. 1 at the position (i,j)(i,j) implies that the ordered pair at the position (i,1)(i,1) of PP is included in the synaptic configuration (j,1)(j,1) of xx. We can think in every column of the matrix as a synaptic configuration. As a consequence of definitions of the synaptic configurations (definition 5-7), a column representing a tripartite synapse has 1 at three positions, a column representing a gap junction has 1 at two positions, and a column representing a one-one-directed synapses has 1 at only one position. Finally, the vector xx is a nn by 1 matrix; each element represents the number of occurrences of a synaptic configuration in AxA_{x} (CountAxCount_{A_{x}}). Note that all the elements of the matrices PP, AA, and xx are always non-negative integers.

There is an inherent challenge in the interpretation of Υ\Upsilon-gm equations due to lack of a natural ordering of the synaptic configurations and their elements (information flows), i.e., any ordered pair could be labeled as α1\alpha_{1} or αn\alpha_{n}, and any synaptic configuration could be labeled as s1s_{1} or sns_{n}. Hence, the same network could be represented in a finite number of ways which is the factorial of the number of synaptic configurations times the factorial of number of information flows. Therefore, when we compare two different networks based on their Υ\Upsilon-gm equations, some counterintuitive situations may arise. We explore these situations in the following.

  1. 1.

    False negative: Given two Υ\Upsilon-gm equations P=AxP=Ax and Q=ByQ=By that represent two neuron-glia networks α\alpha and β\beta. α=β\alpha=\beta does not necessarily imply that P=QP=Q, A=BA=B, or x=yx=y. In other words, different Υ\Upsilon-gm equations can represent the same neuron-glia network.

  2. 2.

    False positive: Given two Υ\Upsilon-gm equations P=AxP=Ax and Q=ByQ=By that represent two neuron-glia networks α\alpha and β\beta. P=QP=Q does not guarantee that A=BA=B or x=yx=y.

  3. 3.

    True comparison: Given two Υ\Upsilon-gm equations P=AxP=Ax and Q=ByQ=By that represent two neuron-glia networks α\alpha and β\beta. If A=BA=B and x=yx=y, then α\alpha and β\beta are isomorphic. (see the last part of the proof of proposition 6)

4 Multidigraphs can misrepresent neuron-glia networks

In this section, we will prove that the existence of isomorphism between the unnested forms of Υ\Upsilon-graphs does not imply that the networks they represent share the same pattern. To do so, we found pairs of non-isomorphic Υ\Upsilon-graphs with isomorphic unnested forms. In the following, we will show four scenarios where isomorphism of the unnested form does not implies isomorphism of the Υ\Upsilon-graphs. Keep in mind that 𝐏\mathbf{P} is a non-empty set of Υ\Upsilon-graphs. At the end of the section (proposition 6), we will prove, in the general case, that the isomorphism between the unnested forms of Υ\Upsilon-graphs implies the isomorphism of their Υ\Upsilon-graphs under some specific conditions.

  1. 1.

    All types of synaptic configuration can be present (proposition 2).

  2. 2.

    There are only tripartite configurations (proposition 3).

  3. 3.

    There are only tripartite configurations and one-one-directed configurations (proposition 4).

  4. 4.

    There are only tripartite configurations, and symmetric configurations (gap junctions) (proposition 5).

To prove propositions 2-5, we use theoretical examples where the isomorphism between the unnested forms does not imply the isomorphism between Υ\Upsilon-graphs. For example, to prove proposition 2, we present two small network motifs (represented by matrices A5×2A_{5\times 2} and B5×3B_{5\times 3}) formed by two neurons and one astrocyte linked by one tripartite synapse (note that matrix AA and BB have 1 as entry of first three rows of the first column, representing a tripartite synapse) together with two one-one-directed synapses (two columns with 1 at one position in AA) or one gap junction(1 column with 1 at two positions in BB). We prove that the unnested form of of these different motifs is isomorphic, meaning that it is possible to find non-isomorphic neuron-glia networks with isomorphic unnested forms. We provided a formal explanation in the in the following proof. In this proof, we establish basic mathematical notation for the following proofs because it is easier to explain them in a simple example rather than in more complicated examples or in the general case (proposition 6). Keep in mind that all examples showed bellow are theoretical examples and when we say “there are”, it means “it is theoretical possible” since the examples do not come from a database.

Proposition 2.

There are two Υ\Upsilon-graphs aa and bb (all synaptic configuration allowed) such that the unnested form of aa and bb are α\alpha-isomorphic (ϕ(a)αϕ(b)\phi(a)\cong_{\alpha}\phi(b)) but aa and bb are non-isomorphic (¬(ab)\neg(a\cong b)).

Proof.

Let’s define the elements of the proposition as follows, a:=(X,Ax)a:=(X,A_{x}), b:=(Y,Ay)b:=(Y,A_{y}), ϕ(a):=(X,preϕ(Ax))\phi(a):=(X,pre-\phi(A_{x})), and ϕ(b):=(Y,preϕ(Ay))\phi(b):=(Y,pre-\phi(A_{y})). Now, since aba\cong b if and only if aa is homomorphic to bb and bb is homomorphic to aa, then we define its negation as

¬(ab)a is non-homomorphic to bb is non-homomorphic to a\neg(a\cong b)\iff a\textit{ is non-homomorphic to }b\lor b\textit{ is non-homomorphic to }a (6)

Since ϕ(a)αϕ(b)\phi(a)\cong_{\alpha}\phi(b) implies that there exists Υ\Upsilon-gm equations P=AxP=Ax and Q=B0yQ=B_{0}y such that they represent the neuron-glia networks of aa and bb, respectively, and P=TQP=TQ (lemma 1, appendix 9.4), we see that TQ=TB0yTQ=TB_{0}y represents the second neuron-glia network. Being B=TB0B=TB_{0}, we will use P=AxP=Ax and P=ByP=By to refer to the Υ\Upsilon-gm equations of the neural networks represented by aa and bb, respectively. Then, using the language of Υ\Upsilon-gm equations, in some cases ABA\neq B or xyx\neq y will imply (6) while holding ϕ(a)αϕ(b)\phi(a)\cong_{\alpha}\phi(b) true. This is the approach that we will follow to prove the first proposition.

Let A and B be the following matrices,

A=A= [1010100101]\begin{bmatrix}1&0\\ 1&0\\ 1&0\\ 0&1\\ 0&1\\ \end{bmatrix} and B=B= [100100100010001]\begin{bmatrix}1&0&0\\ 1&0&0\\ 1&0&0\\ 0&1&0\\ 0&0&1\\ \end{bmatrix}

If x=[α,β]Tx=[\alpha,\beta]^{T} and y=[α,β,β]Ty=[\alpha,\beta,\beta]^{T}, then AxAx and ByBy are equal to P=[α,α,α,β,β]TP=[\alpha,\alpha,\alpha,\beta,\beta]^{T}. We realize that it is impossible to establish aba\cong b as AxA_{x}^{*} and AyA_{y}^{*} have different cardinality. Therefore, we have proved a case in which ϕ(a)αϕ(b)\phi(a)\cong_{\alpha}\phi(b) and ¬(ab)\neg(a\cong b). This case can be easily generalized to networks containing aa and bb as sub-networks.

The previous instance (C1C_{1}) reflects the lack of “resolution power” of models that only use ordered pairs as elements such as digraphs or multidigraphs. Indeed, at a physiological, the neuron-glia networks represented by aa and bb will not show the same behaviour since two one-one-directed synapses are not equivalent to a gap junction. However, due to the relative difficulty to find gap junctions in networks due to technical limitations, we present the following instance of CC that does not posses them. To prove proposition 3, we find that two solutions for an Υ\Upsilon-gm equation (defined in 5), proving that there can be non-isomorphic Υ\Upsilon-graphs with isomorphic unnested forms even when gap junctions are not allowed. We follow the same approach to prove proposition 3 - 5.

Proposition 3.

There are two Υ\Upsilon-graphs aa and bb (only tripartite configurations allowed) such that the unnested form of aa and bb are α\alpha-isomorphic (ϕ(a)αϕ(b)\phi(a)\cong_{\alpha}\phi(b)) but aa and bb are non-isomorphic (¬(ab)\neg(a\cong b)).

Proof.

Using the same notation as in the previous proof, from this case on, we will use a basic concept of linear algebra to prove the existence of Υ\Upsilon-graphs that hold the statements. We can make the coefficient matrices equal (A=BA=B), and then find different vectors that solve P=AxP=Ax. Each solution (vector) corresponds to a different Υ\Upsilon-graph; therefore, after proving that the system has multiple solutions we will show that they represent non-isomorphic Υ\Upsilon-graphs. In this way, we start showing that the null space of AA has more solutions than the zero vector. Let AA be the following matrix,

A=A= [110000001010000010010000010010000100010000100100001000100001100000010010000010010000010100000011]\begin{bmatrix}1&1&0&0&0&0&0&0\\ 1&0&1&0&0&0&0&0\\ 1&0&0&1&0&0&0&0\\ 0&1&0&0&1&0&0&0\\ 0&1&0&0&0&1&0&0\\ 0&0&1&0&0&1&0&0\\ 0&0&1&0&0&0&1&0\\ 0&0&0&1&1&0&0&0\\ 0&0&0&1&0&0&1&0\\ 0&0&0&0&1&0&0&1\\ 0&0&0&0&0&1&0&1\\ 0&0&0&0&0&0&1&1\\ \end{bmatrix}

Ax=0Ax=0 has non-zero solutions such as s=[1,1,1,1,1,1,1,1]Ts=[1,-1,-1,-1,1,1,1,-1]^{T} and any vector ksks where kk is a positive integer. This implies that P=AxP=Ax has either 0 or infinite solutions.

Now we will take a case in point, x1=[3,1,1,1,3,3,3,1]Tx_{1}=[3,1,1,1,3,3,3,1]^{T} and x2=[2,2,2,2,2,2,2,2]Tx_{2}=[2,2,2,2,2,2,2,2]^{T}, we see that Ax1Ax_{1} and Ax2Ax_{2} are equal to the vector P=[4,4,4,4,4,4,4,4]TP=[4,4,4,4,4,4,4,4]^{T}. It is clear that the Υ\Upsilon-graphs that are represented by x1x_{1} and x2x_{2} are not isomorphic since it is impossible to establish an homomorphism between aa and bb since 323\neq 2 and 121\neq 2.
In ordinary words, this means that are neuron-glia networks that share the same CountCount of information flows (elements of the synaptic configurations) while holding different CountCounts of synaptic configurations. Finally, networks containing AA as a part of the coefficient matrix may also satisfy the general statement ϕ(a)αϕ(b)¬(ab)\phi(a)\cong_{\alpha}\phi(b)\land\neg(a\cong b).

Since the previous matrix equation may be a little hard to find in a real network due to combinatory of arrangements between seven tripartite synapses, we present the following motif that is most likely to be part of a neuron-glia network because it has fewer tripartite and some one-one-directed synapses.

Proposition 4.

There are two Υ\Upsilon-graphs aa and bb (symmetric configuration not allowed) such that the unnested form of aa and bb are α\alpha-isomorphic (ϕ(a)αϕ(b)\phi(a)\cong_{\alpha}\phi(b)) but aa and bb are non-isomorphic (¬(ab)\neg(a\cong b)).

Proof.

Continuing with the same notation, likewise the above case, let AA be the following matrix,

A=A= [1100000010001000101000000010001000001010000000110000110000110000]\begin{bmatrix}1&1&0&0&0&0&0&0\\ 1&0&0&0&1&0&0&0\\ 1&0&1&0&0&0&0&0\\ 0&0&1&0&0&0&1&0\\ 0&0&0&0&1&0&1&0\\ 0&0&0&0&0&0&1&1\\ 0&0&0&0&1&1&0&0\\ 0&0&1&1&0&0&0&0\\ \end{bmatrix}

Here, the solutions of Ax=0Ax=0 are of the form ksks, where s=[s=[1, -1, -1, 1, -1, 1, 1, -1]T]^{T} and kk is a positive integer. As an example, let x1=[3,1,1,3,1,3,3,1]Tx_{1}=[3,1,1,3,1,3,3,1]^{T} and x2=[2,2,2,2,2,2,2,2]Tx_{2}=[2,2,2,2,2,2,2,2]^{T}, Ax1Ax_{1} and Ax2Ax_{2} are equal to P=[4,4,4,4,4,4,4,4]TP=[4,4,4,4,4,4,4,4]^{T}. Using the same reasoning as above, the Υ\Upsilon-graphs that are represented by x1x_{1} and x2x_{2} are not isomorphic. ∎

The following case disregards one-one-directed synapses altogether to show that even in such particular scenario our proposition is still true.

Proposition 5.

There are two Υ\Upsilon-graphs aa and bb (one-one-directed configuration not allowed) such that the unnested form of aa and bb are α\alpha-isomorphic (ϕ(a)αϕ(b)\phi(a)\cong_{\alpha}\phi(b)) but aa and bb are non-isomorphic (¬(ab)\neg(a\cong b)).

Proof.

Following the same notation, in this scenario, the coefficient matrix of interest AA is showed in appendix 9.6. The solutions of Ax=0Ax=0 are of the form ksks, where s=[1,1,1,1,1,1,1,1,1,1,1,1,1,1,1,1,1,1,1,1,1,1]Ts=[-1,1,1,-1,-1,1,1,-1,-1,1,1,-1,-1,-1,1,1,1,1,-1,-1,1,-1]^{T} as a solution for Ax=0Ax=0. and kk is a positive integer. As an example, let x1=[1,3,3,1,1,3,3,1,1,3,3,1,1,1,3,3,3,3,1,1,3,1]Tx_{1}=[1,3,3,1,1,3,3,1,1,3,3,1,1,1,3,3,3,3,1,1,3,1]^{T} and x2=[2,2,2,2,2,2,2,2,2,2,2,2,2,2,2,2,2,2,2,2,2,2]Tx_{2}=[2,2,2,2,2,2,2,2,2,2,2,2,2,2,2,2,2,2,2,2,2,2]^{T}, Ax1Ax_{1} and Ax2Ax_{2} are equal to P=[4,4,4,4,4,4,4,4,4,4,4,4,4,4,4,4,4,4,4,4,4]TP=[4,4,4,4,4,4,4,4,4,4,4,4,4,4,4,4,4,4,4,4,4]^{T}. Once again, the Υ\Upsilon-graphs that are represented by x1x_{1} and x2x_{2} are not isomorphic. ∎

In propositions 2-5, we realized that for a pair of representations of a neuron-glia network aa and ϕ(a)\phi(a), where aa is a Υ\Upsilon-graph and ϕ(a)\phi(a), its unnested form, it is safer to work with aa rather than ϕ(a)\phi(a) since it is theoretically possible to lose information during the unnesting of the Υ\Upsilon-graph. It is important to mention that some of the showed examples may not exist in real networks but they prove that it is not safe to assume that the isomorphism between the unnested forms (ϕ(a)αϕ(b)\phi(a)\cong_{\alpha}\phi(b)) implies the isomorphism between Υ\Upsilon-graphs (aba\cong b). However, we conjectured that

ϕ(a)αϕ(b)ab\phi(a)\cong_{\alpha}\phi(b)\implies a\cong b (7)

holds under some specific conditions. The conditions in which 7 is true are defined in the following paragraph and, in proposition 6, we prove that these conditions are sufficient for 7 to be true.

Q1Q_{1} and Q2Q_{2} are conditions applied over two Υ\Upsilon-graphs aa (with AaA_{a} as its Υ\Upsilon-multiset) and bb (with AbA_{b} as its Υ\Upsilon-multiset). Being Ax=PAx=P and B0y=QB_{0}y=Q, the Υ\Upsilon-gm equations related to aa and bb, respectively, if rank(A)rank(A) and rank(B0)rank(B_{0}) are the ranks of AA and B0B_{0}, respectively, then Q1rank(A)=Aarank(B0)=AbQ_{1}\equiv rank(A)=\mid A_{a}^{*}\mid\land\ rank(B_{0})=\mid A_{b}^{*}\mid. This implies that all the columns of AA and B0B_{0} are linearly independent in their matrices. Being TT a row exchange matrix, Q2Q_{2} means that if x=Tyx=Ty, then A=TB0A=TB_{0}.

Proposition 6.

For every pair of Υ\Upsilon-graphs, aa and bb, such that they hold Q1Q_{1} and Q2Q_{2} and the unnested forms of aa and bb (ϕ(a)\phi(a) and ϕ(b)\phi(b)) are α\alpha-isomorphic, then aa and bb are isomorphic.

Proof.

Q1Q_{1} guarantees that both Υ\Upsilon-gm equations have either 1 or no solutions since the null space of the coefficient matrices AA and B0B_{0} are trivial, i.e., only contain the zero vector as a solution. Now, given the definition of ϕ(a)\phi(a) (definition 14), the equations are always solvable, therefore they have 1 and only 1 solution.
Since Lemma 1, P=TQP=TQ where TT is a row exchange matrix. Therefore, we multiply TT to both sides of the Υ\Upsilon-gm equation of bb. The result, TB0y=PTB_{0}y=P, is also a valid Υ\Upsilon-gm equation that represents the relationship between bb and ϕ(b)\phi(b). Let TB0=B1TB_{0}=B_{1}, we have the following equations

{Ax=PB1y=P\begin{cases}Ax=P\\ B_{1}y=P\end{cases}

Now, due to condition Q2Q_{2}, B1=AB_{1}=A; therefore, Ax=AyAx=Ay. Now, realize that since AA has a trivial null space, Ax=AyAx=Ay implies that x=yx=y. This is explained with the following reasoning,

Ax=AyA(x+(1)y)=Ax+A(y)A(xy)=Ay+A(y)=AyAyA(xy)=0xy=0x=y\begin{array}[]{ccl}Ax=Ay&\implies&A(x+(-1)y)=Ax+A(-y)\\ &\implies&A(x-y)=Ay+A(-y)=Ay-Ay\\ &\implies&A(x-y)=0\\ &\implies&x-y=0\\ &\implies&x=y\end{array}

The fourth step is based on the fact that Aα=0A\alpha=0 has only one solution which is α=0\alpha=0. To summarize, Ax=PAx=P is the Υ\Upsilon-gm equation of the Υ\Upsilon-graphs aa and bb. The last step is to show that this implies that aba\cong b. Let’s define both Υ\Upsilon-graphs as a:=(X,Ax)a:=(X,A_{x}) and b:=(Y,Ay)b:=(Y,A_{y}). Keep in mind that aa and bb share the same types (A=AA=A), number (A=AA=A), and repetition of synapses (x=xx=x). Now, let’s choose a function among all the bijections that can be defined between XX and YY such that

a,bX,(a,b)preϕ(Ax)(f(a),f(b))preϕ(Ay)\forall a,b\in X,(a,b)\in pre-\phi(A_{x})\implies(f(a),f(b))\in pre-\phi(A_{y})

We can verify that α:={(a,b):a,bX}Axαf:={(f(a),f(b)):f(a),f(b)Y}Ay\alpha:=\{(a,b):a,b\in X\}\in A_{x}\implies\alpha_{f}:=\{(f(a),f(b)):f(a),f(b)\in Y\}\in A_{y} since every ordered pair from preϕ(Ax)pre-\phi(A_{x}) is 1-1 associated with a pair of preϕ(Ay)pre-\phi(A_{y}). However, since aa and bb share the same coefficient matrix (AA) and CountCount vector (xx), the following is true for every synapse of AxA_{x}, CountAx(α)=CountAy(αf)Count_{A_{x}}(\alpha)=Count_{A_{y}}(\alpha_{f}), in other words, ff is an homomorphism from XX to YY. We can analogously define an homomorphism from YY to XX, concluding that aba\cong b.

5 Modeling the network shape

Nowadays, a huge amount of information about neural shape and network connectivity is available in open databases [7, 38, 41]. However, the mathematical tools to study neuronal shape are still under development [19, 22]. Natural structure theory could contribute to this research field by providing a theoretical framework to study the shape of neurons, astrocytes, or even networks [3, 4]. In this section, we use a mathematical formalism (definition 15) that represents the concept of a natural structure [3, 4] to model the morphology of a neuron-glia network.

A connected neuron-glia network is evidently a natural structure, where its elements are neural-like cells and single synapses are NE-connections [3, 4]. In general, a single synapse is composed of a presynaptic membrane, an astrocyte membrane (if it is tripartite), a postsynaptic membrane, and a perisynaptic matrix (if it is chemical). The synaptic cleft is a region of the space filled of perisynaptic matrix which is composed of proteoglycans and proteins [25, 17, 27]. This matrix mediates trans-synaptic adhesion (NE-connection role) and regulates the synaptic transmission (communication role) [17, 40]. Thus, a single synapse can be thought of as an structure. In this section, we roughly model a neuron-glia network, applying part of the mathematical definition of natural structures (definition 15) from Cabrera-Febola, 2023-in preparation for publication.

Definition 15.

(Cabrera-Febola, 2023-in preparation for publication) A natural structure (NE) is a triple (X,D,C)(X,D,C), where XX is a collection of the elements of the NE (X\varnothing\notin X), CC is a family of collections of NE-connections, and DD is a distribution of the elements, being a pair (SS,{msS:mX,s=1,2,,v,v}\{m_{s}\hookrightarrow S:m\in X,s=1,2,...,v,v\in\mathbb{N}\}), where SS is the “known” natural space and msSm_{s}\hookrightarrow S stands for msm_{s} exists along SS.

In our case, the natural structure model of a connected neuron-glia network (connected Υ\Upsilon-graph) whose synaptic collection is represented by a Υ\Upsilon-family F:ISXF:I\mapsto S_{X} is (X,𝔻,I)(X,\mathbb{D},I) where XX is a set of neuron-like cells (astrocytes and neurons) and II is a set of single synapses (NE-connections). In order to define 𝔻\mathbb{D} which models the distribution of elements, we define a distribution function g:XIPg:X\cup I\to P where PP is a family of subsets of 3\mathbb{R}^{3} such that any element xx of the domain of gg is distributed in all the points of g(x)g(x) and nowhere else. Hence, we can define 𝔻\mathbb{D} as (3,g)(\mathbb{R}^{3},g) to model the spatial distribution of the network.

Neuron-like cells have not cavities, so we can assume that g(x)g(x) with xXx\in X is homeomorphic to a solid ball. On the other hand, g(s)g(s) with sCs\in C is the space where neurotransmission occurs, thus it can be assumed to be also homeomorphic to a solid ball. Additionally, we can state that synapses are partially composed of membranes from all implied neuron-like cells, so g(s)g(xi)g(xj)g(xz)g(s)\cap g(x_{i})\cap g(x_{j})\cap...\cap g(x_{z})\neq\varnothing if and only if sIs\in I and F(s)F(s) is a relation on {xi,xj,,xz}\{x_{i},x_{j},...,x_{z}\}. Therefore, since the Υ\Upsilon-graph is connected, it is possible to conjecture that the {g(e):eXI}\{\bigcup g(e):e\in X\cup I\} is a solid. There should be more topological and geometrical properties of neuron-glial networks; however, this topic is out of the scope of our article.

6 Discussion

Neurons has been the focus of two hundreds of years of neuroscience research [33]. However, experimental evidence suggests that astrocytes play an important role in neuronal dynamics, making it necessary to deepen our understanding of how astrocytes network influences neural phenomena [6, 10, 21]. Although there were some attempts to model neuron-glia network dynamics, they lack of a formal network representation [5, 34, 39]. To fill this lack, in this study, we present a formal representations of a neuron-glia network, the Υ\Upsilon-graph.

An Υ\Upsilon-graph is defined as a tuple (X,SX)(X,S_{X}) where XX is a set of neural-like cells and SXS_{X} a multiset of synaptic configurations. In our framework, synaptic configurations are sets of ordered pairs that represent the information flux inside a synapse. Since neurons can establish many synapses with the same information flux, i.e. the same synaptic configuration, SXS_{X} is a multiset whose counting function allows us to know the number of occurrence of a synaptic configurations in the network.

Previous works on neural networks had not considered each synapse as an individual element but the collection of synapses as the edge of the digraph or graph [26, 28, 35]. However, graph-like representations can misrepresent a neuronal network. For example, if a neuronal network has electrical synapses, each electrical synapse is equivalent to two chemical synapses with inverse directions in a digraph. In contrast, our work presents a new mathematical object, Υ\Upsilon-graph which consider each single synapses (electrical, chemical, or tripartite synapse) as an element of the network.

An alternative representation that would address the issue of the multiple types of synapses in the brain would be a multilayer network [9]. In this mathematical object, a layer can represent a type of synapses, solving the misrepresentation problem discussed in the previous paragraph. However, multilayer networks like digraphs assign arcs between two vertices, thus misrepresenting the triadic nature of tripartite synapses. This misrepresentation is an issue when two different networks are represented by the same mathematical objects. In proposition 3, we found a theoretical example of multidigraph that can represent multiple different neuron-glia networks with only tripartite synapses. This misrepresentation can not be fixed by using multilayer networks because there is only one type of synapse in our example.

In addition to Υ\Upsilon-graphs, there are other mathematical objects that can represent the information flow in triadic relations: directed hypergraphs [12] and abstract directed simplicial complex [26]. The problem with directed hypergraphs is that a vertex can not be part of the input and outputs of an hyperarc, meaning that, in our context, a neuron-like cell (i.e. astrocyte or neuron) could only send information but not receive information throughout the same synapse. This is not the case for tripartite synapses because the astrocyte receives information from the presynaptic neuron and sends information to the postsynaptic neurons [15, 29, 36, 18]. The case of simplicial complexes need further study to prove wether it is always possible to built a simplicial complex from a Υ\Upsilon-graph or not. Simplicial complexes are used to study the algebraic topology of networks [16, 26, 31, 32] while graph-like objects are used to find network motifs and to make graph measures such as between centrality and modularity [2].

Since graph representations are widely used in neuroscience, we defined a function (preϕpre-\phi, see appendix 9.2) that unnests Υ\Upsilon-graphs into multidigraphs. However, using linear algebra, we proved that the isomorphism between unnested forms does not guarantee isomorphism between their respective Υ\Upsilon-graphs in several scenarios (propositions 2-5). Therefore, multidigraph representations can misrepresent a neuron-glia networks. Interestingly, we also found that if the coefficient matrices and the vectors of synaptic configurations of the two Υ\Upsilon-graphs are equal after interchanging rows, and the coefficient equations’ columns are linearly independent, the isomorphism between unnested forms guarantees the isomorphism between Υ\Upsilon-graphs. Under these specific conditions, we can use graph theoretical tools developed for multidigraphs without misrepresentation issues.

In this work, we also sketched a novel approach to model the network shape. As Υ\Upsilon-family represents each single synapse as a distinguishable element, we use this object to create a natural structure model. This natural structure model is a tuple (X,𝔻,I)(X,\mathbb{D},I) where XX is a set of neural-like cells, II is a set of single synapses,and 𝔻\mathbb{D} is a tuple (3,g)(\mathbb{R}^{3},g) where gg is the distribution function. In our model, single synapses and neuron-like cells are homeomorphic to solid balls. We conjecture that the portion of space where neuron-like cells and synapses are distributed is also a solid. This hypothetical solid could be studied applying differential geometry or algebraic topology which open new ways for studying structural connectomes.

In conclusion, we have developed a mathematical framework for working with connectivity structure of neuron-glia networks (Υ\Upsilon-graphs, conectedness, isomorphism, and other constructs.) We have also connected the concept of Υ\Upsilon-graphs with graph-like objects, multidigraphs, and studied their relations (ϕ\phi function and matrix equation). On the other hand, we have sketched a natural structure model to represent the spatial structure of neuron-glia networks.

7 Acknowledgments

We thank D. A. Pacheco (Harvard University) and Leonardo Torres (Max Planck Institute for Mathematics in the Sciences) for their helpful comments and partial manuscript proofreading.

8 Declarations

8.1 CRediT author statement

Marco Peña-Garcia: Conceptualization, Methodology, Investigation, Writing - Original Draft. Francesco Peña-Garcia: Methodology, Investigation, Writing - Original Draft. Nelson Castro: Conceptualization, Methodology, Writing - Review & Editing, Supervision.                                                                                        Walter Cabrera-Febola: Writing - Review & Editing, Supervision.

8.2 Competing interests

The authors have no competing interests that are relevant to the content of this article.

9 Supplementary section

9.1 One NL-set per Υ\Upsilon-multiset

In this appendix, we prove that there is only one NL-set per Υ\Upsilon-msets. First, lets define a relation \diamond between Υ\Upsilon-msets and NL-sets as follows

αβ(α𝒫(β×β)(xβyα:xDom(y)xRan(y)))\alpha\diamond\beta\iff(\alpha^{*}\subset\mathcal{P}(\beta\times\beta)\land(\forall x\in\beta\exists y\in\alpha:x\in Dom(y)\lor x\in Ran(y)))

Now, let 𝐌\mathbf{M} be an arbitrary non-empty set of Υ\Upsilon-msets and 𝐋\mathbf{L}, an arbitrary non-empty set of NL-sets, then the following is true.

Proposition 7.
A𝐌X,Y𝐋,((AXAY)X=Y)\forall A\in\mathbf{M}\forall X,Y\in\mathbf{L},((A\diamond X\land A\diamond Y)\implies X=Y)
Proof.

We will proceed with a reductio ad absurdum argument. Let PP be the proposition, then

¬PA𝐌X,Y𝐋,(AXAY)XY\neg P\equiv\exists A\in\mathbf{M}\exists X,Y\in\mathbf{L},(A\diamond X\land A\diamond Y)\land X\neq Y
¬P((xX:xY)(yY:yX))((αA:(xDom(α)xRan(α))α𝒫(Y×Y))(βA:(yDom(β)yRan(β))β𝒫(X×X)))¬(AY)¬(AX)¬(AXAY)¬((AXAY)XY)\begin{array}[]{ccc}\neg P&\implies&((\exists x\in X:x\not\in Y)\lor(\exists y\in Y:y\not\in X))\\ &\implies&((\exists\alpha\in A^{*}:(x\in Dom(\alpha)\lor x\in Ran(\alpha))\land\alpha\not\in\mathcal{P}(Y\times Y))\lor\\ &&(\exists\beta\in A^{*}:(y\in Dom(\beta)\lor y\in Ran(\beta))\land\beta\not\in\mathcal{P}(X\times X)))\\ &\implies&\neg(A\diamond Y)\lor\neg(A\diamond X)\\ &\implies&\neg(A\diamond X\land A\diamond Y)\\ &\implies&\neg((A\diamond X\land A\diamond Y)\land X\neq Y)\\ \end{array}

()(\Rightarrow\!\Leftarrow)

A𝐌X,Y𝐋,((AXAY)X=Y)\therefore\forall A\in\mathbf{M}\forall X,Y\in\mathbf{L},((A\diamond X\land A\diamond Y)\implies X=Y)

You can also understand this proof as “there is no AA, XX, and YY such that ¬P\neg P is true, it follows that if 𝐌\mathbf{M} and 𝐋\mathbf{L} are non-empty, then they hold PP”. Remember that ¬(¬P)P\neg(\neg P)\equiv P.

9.2 Defining preϕpre-\phi function

In this appendix, we formally define the preϕpre-\phi function in a constructive way as follows.

Let 𝐀\mathbf{A} be an arbitrarily large set of finite multisets and 𝐀m\mathbf{A}^{m} an arbitrarily large set of sets. For any A𝐀A\in\mathbf{A} and B𝐀mB\in\mathbf{A}^{m}, we say ABA\sim B if and only if the following holds.

  1. 1.

    aB,aA×\forall a\in B,a\in A^{*}\times\mathbb{N}.

  2. 2.

    bAaB:{b}a\forall b\in A^{*}\exists a\in B:\{b\}\in a.

  3. 3.

    aB,{c}a{c,CountA(c)}a\forall a\in B,\{c\}\in a\implies\{c,\text{Count}_{A}(c)\}\in a.

Being cc any element, i.e., a free variable of the statement. The first condition means that BB is a set of ordered pairs (b,n)(b,n) where bb is an element of AA^{*} and nn, a natural number. The second condition implies that every element of AA^{*} is represented in BB. The final condition defines nn for every (b,n)(b,n) as the CountCount of bb in AA. Since every bAb\in A^{*} has only one CountCount, every bb is associated with one and only on (b,n)(b,n) in BB.

Since A,B𝐀,C𝐀m,(ABACB=C)\forall A,B\in\mathbf{A},C\in\mathbf{A}^{m},(A\sim B\land A\sim C\implies B=C), and 𝐀\mathbf{A} and 𝐀m\mathbf{A}^{m} are sets, the relation \sim defines a function from 𝐀\mathbf{A} to 𝐀m\mathbf{A}^{m}. We denote this function as π\pi.

π:𝐀𝐀mAB=π(A)\begin{array}[]{rl}\pi:\mathbf{A}\to&\mathbf{A}^{m}\\ A\mapsto&B=\pi(A)\end{array}

where B=π(A)ABB=\pi(A)\iff A\sim B. It is clear that π\pi is injective.
There exists a subset of π(𝐀)\pi(\mathbf{A}) whose elements has the following form,

B:={(α,n):αSXn}B:=\{(\alpha,n):\alpha\in S_{X}\land n\in\mathbb{N}\}

where SXS_{X} is a set of binary relations on subsets of some NL-set XX. This subset of π(𝐀)\pi(\mathbf{A}) will be denoted as Sππ(𝐀)S^{\pi}\subset\pi(\mathbf{A}).

Let 𝕌\mathbb{U} be an arbitrary large set of sets of sets, we define a function δ\delta from SπS^{\pi} to 𝕌\mathbb{U}, i.e.,

δ:Sπ𝕌BC=δ(sπ)\begin{array}[]{rl}\delta:S^{\pi}\to&\mathbb{U}\\ B\mapsto&C=\delta(s^{\pi})\end{array}

δ\delta is defined as

C=δ(B)b1b2C=\delta(B)\iff b_{1}\land b_{2}

where

  1. b1b_{1}\equiv

    (α,n)BcC:(aα(a,n)c)\forall(\alpha,n)\in B\exists c\in C:(a\in\alpha\implies(a,n)\in c).

  2. b2b_{2}\equiv

    |B|=|C||B|=|C|.

Keep in mind that B=π(A)B=\pi(A) for some A𝐀A\in\mathbf{A}, i.e., C=δ(π(A))C=\delta(\pi(A)).
For the next step, we will define a sum operator +c+^{c} as

+c:π(𝐀)×π(𝐀)π(𝐀)(c1,c2)c3=c1+cc2\begin{array}[]{rl}+^{c}:\pi(\mathbf{A})\times\pi(\mathbf{A})\to&\pi(\mathbf{A})\\ (c_{1},c_{2})\mapsto&c_{3}=c_{1}+^{c}c_{2}\end{array} (8)

This operation is defined as

c3=c1+cc2d1d2d3d4c_{3}=c_{1}+^{c}c_{2}\iff d_{1}\land d_{2}\land d_{3}\land d_{4}

where

  1. d1d_{1}

    (a,n)c1(a,m)c2(a,m+n)c3(a,n)\in c_{1}\land(a,m)\in c_{2}\implies(a,m+n)\in c_{3}.

  2. d2d_{2}

    m,(a,n)c1(a,m)c2(a,n)c3\forall m\in\mathbb{N},(a,n)\in c_{1}\land(a,m)\not\in c_{2}\implies(a,n)\in c_{3}.

  3. d3d_{3}

    m,(a,n)c2(a,m)c1(a,n)c3\forall m\in\mathbb{N},(a,n)\in c_{2}\land(a,m)\not\in c_{1}\implies(a,n)\in c_{3}.

  4. d4d_{4}

    If m1:={a:(a,n)c1}m_{1}:=\{a:(a,n)\in c_{1}\} and m2:={a:(a,n)c2}m_{2}:=\{a:(a,n)\in c_{2}\},
    then m1m2=c3\mid m_{1}\cup m_{2}\mid=\mid c_{3}\mid.

In d1d_{1}, m+nm+n refers to the usual sum of integers; in d4d_{4}, nn is a free variable.
With this operator, we define i=1n(ci)\sum_{i=1}^{n}(c_{i}) as the successive application of +c+^{c} to the elements of the sequence (c)i=1n(c)_{i=1}^{n}.
As an example, i=14(ci)=((c1+cc2)+cc3)+cc4\sum_{i=1}^{4}(c_{i})=((c_{1}+^{c}c_{2})+^{c}c_{3})+^{c}c_{4}. With this “series”, we define the function γ\gamma, being δ(Sπ)𝕌\delta(S^{\pi})\subset\mathbb{U} the range of δ\delta,

γ:δ(Sπ)π(𝐀)CD=γ(C)\begin{array}[]{rl}\gamma:\delta(S^{\pi})\to&\pi(\mathbf{A})\\ C\mapsto&D=\gamma(C)\end{array}

Where D=γ(C)D=\gamma(C) if and only if, being (c)i=1n(c)_{i=1}^{n} a sequence of all the elements of CC, D=i=1n(ci)D=\sum_{i=1}^{n}(c_{i}).
The last step consists in transforming DD back into a multiset, we do this with our last function π\pi^{-}.

π:π(𝐀)𝐀DE=π(D)\begin{array}[]{rl}\pi^{-}:\pi(\mathbf{A})\to&\mathbf{A}\\ D\mapsto&E=\pi^{-}(D)\end{array}

π\pi^{-} is defined as

E=π(D)EDE=\pi^{-}(D)\iff E\sim D

Similar to π\pi, this function is also injective. Furthermore, π\pi^{-} is also surjective since its range is the whole codomain 𝐀\mathbf{A}.
Finally, we resume all of these operations with the function preϕpre-\phi,

preϕ:𝐀𝐀AE=preϕ(A)\begin{array}[]{rl}pre-\phi:\mathbf{A}\to&\mathbf{A}\\ A\mapsto&E=pre-\phi(A)\end{array}

preϕpre-\phi is defined as

E=preϕ(A)E=π(γ(δ(π(A))))E=pre-\phi(A)\iff E=\pi^{-}(\gamma(\delta(\pi(A))))

9.3 One NL-set per image of preϕpre-\phi

Given preϕpre-\phi definition, we can prove that there is only one NL-set per image of preϕpre-\phi. Let \diamond^{{}^{\prime}} be a relation between msets of ordered pairs and NL-sets as follows

αβ(α𝒫(β×β)(xβ(y1,y2)α:x=y1x=y2))\alpha\diamond^{{}^{\prime}}\beta\iff(\alpha^{*}\subset\mathcal{P}(\beta\times\beta)\land(\forall x\in\beta\exists(y_{1},y_{2})\in\alpha:x=y_{1}\lor x=y_{2}))

Let 𝐄\mathbf{E} be a non-empty collection of images of the function preϕpre-\phi and 𝐋\mathbf{L}, an arbitrary non-empty set of NL-sets as previously, then the following proposition is true.

Proposition 8.
A𝐄X,Y𝐋,((AXAY)X=Y)\forall A\in\mathbf{E}\forall X,Y\in\mathbf{L},((A\diamond^{{}^{\prime}}X\land A\diamond^{{}^{\prime}}Y)\implies X=Y)
Proof.

Let PP be the proposition, then

¬PA𝐄X,Y𝐋,(AXAY)XY\neg P\equiv\exists A\in\mathbf{E}\exists X,Y\in\mathbf{L},(A\diamond^{{}^{\prime}}X\land A\diamond^{{}^{\prime}}Y)\land X\neq Y
¬P((xX:xY)(yY:yX))(((α1,α2)A:α1Yα2Y)(((β1,β2)A:β1Xβ2X))(((α1,α2)A:(α1,α2)A)((β1,β2)A:(β1,β2)A))AA\begin{array}[]{ccc}\neg P&\implies&((\exists x\in X:x\not\in Y)\lor(\exists y\in Y:y\not\in X))\\ &\implies&((\exists(\alpha_{1},\alpha_{2})\in A:\alpha_{1}\not\in Y\lor\alpha_{2}\not\in Y)\lor\\ &&((\exists(\beta_{1},\beta_{2})\in A:\beta_{1}\not\in X\lor\beta_{2}\not\in X))\\ &\implies&((\exists(\alpha_{1},\alpha_{2})\in A:(\alpha_{1},\alpha_{2})\not\in A)\lor\\ &&(\exists(\beta_{1},\beta_{2})\in A:(\beta_{1},\beta_{2})\not\in A))\\ &\implies&A\neq A\end{array}

()(\Rightarrow\!\Leftarrow)

A𝐄X,Y𝐋,((AX(AY)X=Y)\therefore\forall A\in\mathbf{E}\forall X,Y\in\mathbf{L},((A\diamond^{{}^{\prime}}X\land(A\diamond^{{}^{\prime}}Y)\implies X=Y)

9.4 Lemma for unnested forms

Lemma 1.

Let ϕ(a)=(X,preϕ(Aa))\phi(a)=(X,pre-\phi(A_{a})) and ϕ(b)=(Y,preϕ(Ab))\phi(b)=(Y,pre-\phi(A_{b})) be multidigraphs obtained by applying ϕ\phi to the Υ\Upsilon-graphs aa and bb, respectively. Let Pa=(ci1)n×1P_{a}=(c_{i1})_{n\times 1} be a column matrix where every element ci1=Countpreϕ(Aa)(αi)c_{i1}=Count_{pre-\phi(A_{a})}(\alpha_{i}), αipreϕ(Aa)\alpha_{i}\in pre-\phi(A_{a})^{*} with i[1,n]i\in[1,n]\cap\mathbb{Z}, being nn the cardinality of preϕ(Aa)pre-\phi(A_{a})^{*}. Analogously for bb, let Pb=(dj1)m×1P_{b}=(d_{j1})_{m\times 1} be a column matrix where every element dj1=Countpreϕ(Ab)(βj)d_{j1}=Count_{pre-\phi(A_{b})}(\beta_{j}), βjpreϕ(Ab)\beta_{j}\in pre-\phi(A_{b})^{*} with j[1,m]j\in[1,m]\cap\mathbb{Z}, being mm the cardinality of preϕ(Ab)pre-\phi(A_{b})^{*}. Then the following is true.

a,b𝐏,ϕ(a)αϕ(b)(T:TPa=Pb)\forall a,b\in\mathbf{P},\phi(a)\cong_{\alpha}\phi(b)\implies(\exists T:TP_{a}=P_{b})

In other words, the elements of PaP_{a} can be rearranged into PbP_{b}. This does not imply that Pa=PbP_{a}=P_{b} although it certainly does not exclude it (T=In×nT=I_{n\times n}).

Proof.

The first thing to realize is that if ϕ(a)αϕ(b)\phi(a)\cong_{\alpha}\phi(b), then there exists a bijection ff between XX and YY such that

Countpreϕ(Aa)(p,q)=Countpreϕ(Ab)(f(p),f(q))Count_{pre-\phi(A_{a})}(p,q)=Count_{pre-\phi(A_{b})}(f(p),f(q))

for any p,qpreϕ(Aa)p,q\in pre-\phi(A_{a}). Since for any (p,q)preϕ(Aa)(p,q)\in pre-\phi(A_{a})^{*} there exists a unique element (f(p),f(q))preϕ(Ab)(f(p),f(q))\in pre-\phi(A_{b})^{*} and viceversa, we can define a function gg which is also a bijection; however, this function will be defined from preϕ(Aa)pre-\phi(A_{a})^{*} to preϕ(Ab)pre-\phi(A_{b})^{*} which, in turn, implies that preϕ(Aa)pre-\phi(A_{a})^{*} and preϕ(Aa)pre-\phi(A_{a})^{*} have the same cardinality. With this, we can affirm that the column matrices PaP_{a} and PbP_{b} have the same number of elements, i.e., n=mn=m.
Now, since every element ci1c_{i1} of PaP_{a} is associated with one and only one element of preϕ(Aa)pre-\phi(A_{a})^{*} and the same for the elements dj1d_{j1} of PbP_{b} with preϕ(Ab)pre-\phi(A_{b})^{*}, then, if PaP_{a}^{\star} and PbP_{b}^{\star} are the sets that contain the elements of PaP_{a} and PbP_{b}, respectively, we can define the following function

h:PaPbci1dj1\begin{array}[]{cccc}h:&P_{a}^{\star}&\to&P_{b}^{\star}\\ &c_{i1}&\mapsto&d_{j1}\end{array}

with dj1=h(ci1)d_{j1}=h(c_{i1}) if and only if (cj1=Countpreϕ(Aa)(α)dj1=Countpreϕ(Ab)(g(α)))(c_{j1}=Count_{pre-\phi(A_{a})}(\alpha)\implies d_{j1}=Count_{pre-\phi(A_{b})}(g(\alpha))). Since gg is a bijection, it is clear that hh is also a bijection. This implies that every element dj1d_{j1} can be defined as h(ci1)h(c_{i1}), in other words, Pa=(ci1)n×1P_{a}=(c_{i1})_{n\times 1} and Pb=(h(ci1))n×1P_{b}=(h(c_{i1}))_{n\times 1}. Now, by the definition of gg and the fact that ϕ(a)αϕ(b)\phi(a)\cong_{\alpha}\phi(b), we say that ci1Pa,h(ci1)=ci1\forall c_{i1}\in P_{a}^{\star},h(c_{i1})=c_{i1}. Together with the new definition of PbP_{b}, this implies that Pa=PbP_{a}^{\star}=P_{b}^{\star}. Therefore PbP_{b} can only be equal to TPaTP_{a} where TT is a row exchange matrix. We complete this proof by providing such TT.
Let T=(αij)T=(\alpha_{ij}), where

αij={1,cj1=di10,cj1di1\alpha_{ij}=\begin{cases}1,&c_{j1}=d_{i1}\\ 0,&c_{j1}\neq d_{i1}\end{cases}

where cc represents an element of PaP_{a} and dd, an element of PbP_{b}. ∎

9.5 Proof of proposition 1

Let αpreϕ(Ax)\alpha\in pre-\phi(A_{x}) be any vertex of ϕ(x)\phi(x) (i.e., an ordered pair), if s1,s2,,sks_{1},s_{2},...,s_{k} are all synaptic configurations (elements of AxA_{x}) that include α\alpha (αsi\alpha\in s_{i} for all i[1,k]i\in[1,k]\cap\mathbb{Z}), then

Countpreϕ(Ax)(α)=CountAx(s1)+CountAx(s2)++CountAx(sk)Count_{pre-\phi(A_{x})}(\alpha)=Count_{A_{x}}(s_{1})+Count_{A_{x}}(s_{2})+...+Count_{A_{x}}(s_{k})

Proof.

From the definition of preϕpre-\phi (appendix 9.2), (α,Countpreϕ(Ax)(α))(\alpha,Count_{pre-\phi(A_{x})}(\alpha)) is an element of D=γ(δ(π(Ax)))D=\gamma(\delta(\pi(A_{x}))). However, if s1,,sks_{1},...,s_{k} are the representations of the synaptic configurations that include α\alpha, then

(α,Countpreϕ(Ax)(α))=(α,CountAx(s1))+c+c(α,CountAx(sk))(\alpha,Count_{pre-\phi(A_{x})}(\alpha))=(\alpha,Count_{A_{x}}(s_{1}))+^{c}...+^{c}(\alpha,Count_{A_{x}}(s_{k}))

since condition d1d_{1} of the definition of +c+^{c} (defined in 8). In other words, Countpreϕ(Ax)(α)=CountAx(s1)++CountAx(sk)Count_{pre-\phi(A_{x})}(\alpha)=Count_{A_{x}}(s_{1})+...+Count_{A_{x}}(s_{k}). ∎

9.6 Matrix for proposition 5

A=A= [000000000000000000101000000000000000000001100000000000000000010001000000000000000001100000000000000000001001000000000000000000100001000000000000000100010000000000000000100010000000000000000100010000000000000000011000000000000000000010010000000000000000001000100000000000000011000000000000000000010100000000000000000100000001000000000000010100000000000000000000100000100000000000000010100000000000000000100000010000000000000100000010000000000000101000000000000000000010000010000000000000010100000000000000000001000001000000000000010010000000000000000001010000000000000000001100000000000000000000101000000000000000000000100000100000000000000010000100000000000000]\setcounter{MaxMatrixCols}{22}\begin{bmatrix}0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&1&0&1&0\\ 0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&1&1&0\\ 0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&1&0&0&0&1\\ 0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&1&1&0&0&0\\ 0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&1&0&0&1&0&0\\ 0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&1&0&0&0&0&1\\ 0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&1&0&0&0&1&0&0\\ 0&0&0&0&0&0&0&0&0&0&0&0&0&0&1&0&0&0&1&0&0&0\\ 0&0&0&0&0&0&0&0&0&0&0&0&0&1&0&0&0&1&0&0&0&0\\ 0&0&0&0&0&0&0&0&0&0&0&0&0&1&1&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0&0&0&0&0&1&0&0&1&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0&0&0&0&0&1&0&0&0&1&0&0&0&0&0\\ 0&0&0&0&0&0&0&0&0&0&1&1&0&0&0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0&0&1&0&1&0&0&0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&1&0&0&0&0&0&0&0&1&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&1&0&1&0&0&0&0&0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0&1&0&0&0&0&0&1&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0&1&0&1&0&0&0&0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&1&0&0&0&0&0&0&1&0&0&0&0&0&0&0&0\\ 0&0&0&0&0&1&0&0&0&0&0&0&1&0&0&0&0&0&0&0&0&0\\ 0&0&0&0&1&0&1&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0\\ 0&0&0&0&1&0&0&0&0&0&1&0&0&0&0&0&0&0&0&0&0&0\\ 0&0&0&1&0&1&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0\\ 0&0&0&1&0&0&0&0&0&1&0&0&0&0&0&0&0&0&0&0&0&0\\ 0&1&0&0&1&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0\\ 0&1&0&1&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0\\ 1&1&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0\\ 1&0&1&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0&0\\ 0&0&1&0&0&0&0&0&1&0&0&0&0&0&0&0&0&0&0&0&0&0\\ 0&0&1&0&0&0&0&1&0&0&0&0&0&0&0&0&0&0&0&0&0&0\\ \end{bmatrix}

References

  • [1] Jørgen Bang-Jensen and Gregory Z. Gutin “Digraphs: theory, algorithms and applications” The Campus, 4 Crinan Street N1 9XW, London, United kingdom: Springer London, 2009
  • [2] Danielle S Bassett and Olaf Sporns “Network neuroscience” In Nature neuroscience 20.3 Nature Publishing Group, 2017, pp. 353–364
  • [3] Walter Cabrera-Febola “On natural structures: Complementary issues” In Spacetime & Substance 5, 2004, pp. 65–71
  • [4] Walter Cabrera-Febola “On natural structures: The unification of nature” In Spacetime & Substance 5, 2004, pp. 34–41
  • [5] Siow-Cheng Chan, Siew-Ying Mok, Danny Wee-Kiat Ng and Sing-Yau Goh “The role of neuron-glia interactions in the emergence of ultra-slow oscillations” In Biological cybernetics 111.5-6 Springer, 2017, pp. 459–472
  • [6] Oana Chever et al. “Astroglial networks promote neuronal coordination” In Science signaling 9.410 American Association for the Advancement of Science, 2016, pp. ra6–ra6
  • [7] Jody Clements et al. “neuPrint: Analysis Tools for EM connectomics” In BioRxiv Cold Spring Harbor Laboratory, 2020
  • [8] Ana Covelo and Alfonso Araque “Neuronal activity determines distinct gliotransmitter release from a single astrocyte” In Elife 7 eLife Sciences Publications Limited, 2018, pp. e32237
  • [9] Manlio De Domenico “More is different in real-world multilayer networks” In Nature Physics 19.9 Nature Publishing Group UK London, 2023, pp. 1247–1262
  • [10] Tara Deemyad, Joel Lüthi and Nelson Spruston “Astrocytes integrate and drive action potential firing in inhibitory subnetworks” In Nature communications 9.1 Nature Publishing Group, 2018, pp. 1–13
  • [11] Yuki Fujii, Shohei Maekawa and Mitsuhiro Morita “Astrocyte calcium waves propagate proximally by gap junction and distally by extracellular diffusion of ATP released from volume-regulated anion channels” In Scientific reports 7.1 Nature Publishing Group, 2017, pp. 1–15
  • [12] Giorgio Gallo, Giustino Longo, Stefano Pallottino and Sang Nguyen “Directed hypergraphs and applications” In Discrete applied mathematics 42.2-3 Elsevier, 1993, pp. 177–201
  • [13] K.. Girish and John Sunil Jacob “On multiset topologies” In Theory and Applications of Mathematics & Computer Science 2.1, 2012, pp. 37–52
  • [14] K.. Girish and John Sunil Jacob “Relations and functions in multiset context” In Information Sciences 179.6 Elsevier, 2009, pp. 758–768
  • [15] Candela González-Arias and Gertrudis Perea “Gliotransmission at Tripartite Synapses” In Computational Glioscience One New York Plaza, Suite 4600, New York, USA: Springer, 2019, pp. 213–226
  • [16] Bradford R Greening Jr, Noa Pinter-Wollman and Nina H Fefferman “Higher-order interactions: understanding the knowledge capacity of social groups using simplicial sets” In Current Zoology 61.1 Oxford University Press Oxford, UK, 2015, pp. 114–127
  • [17] Seil Jang, Hyejin Lee and Eunjoon Kim “Synaptic adhesion molecules and excitatory synaptic transmission” In Current opinion in neurobiology 45 Elsevier, 2017, pp. 45–50
  • [18] Alistair Jennings et al. “Dopamine elevates and lowers astroglial Ca2+ through distinct pathways depending on local synaptic circuitry” In Glia 65.3 Wiley Online Library, 2017, pp. 447–459
  • [19] Lida Kanari et al. “A topological representation of branching neuronal morphologies” In Neuroinformatics 16.1 Springer, 2018, pp. 3–13
  • [20] Waldemar Karwowski, Farzad Vasheghani Farahani and Nichole Lighthall “Application of graph theory for identifying connectivity patterns in human brain networks: a systematic review” In frontiers in Neuroscience 13 Frontiers, 2019, pp. 585
  • [21] Hosuk Sean Lee et al. “Astrocytes contribute to gamma oscillations and recognition memory” In Proceedings of the National Academy of Sciences 111.32 National Acad Sciences, 2014, pp. E3343–E3352
  • [22] Yanjie Li et al. “Metrics for comparing neuronal tree shapes based on persistent homology” In PloS one 12.8 Public Library of Science San Francisco, CA USA, 2017, pp. e0182184
  • [23] Albert Lin et al. “Network statistics of the whole-brain connectome of Drosophila” In Nature 634.8032 Nature Publishing Group UK London, 2024, pp. 153–165
  • [24] James I Nagy, Alberto E Pereda and John E Rash “Electrical synapses in mammalian CNS: Past eras, present focus and future directions” In Biochimica et Biophysica Acta (BBA)-Biomembranes 1860.1 Elsevier, 2018, pp. 102–123
  • [25] Anna Pintér et al. “Chondroitin sulfate proteoglycan-5 forms perisynaptic matrix assemblies in the adult rat cortex” In Cellular Signalling 74 Elsevier, 2020, pp. 109710
  • [26] Michael W Reimann et al. “Cliques of neurons bound into cavities provide a missing link between structure and function” In Frontiers in computational neuroscience 11 Frontiers, 2017, pp. 48
  • [27] Gabrielle Rudenko “Neurexins—versatile molecular platforms in the synaptic cleft” In Current opinion in structural biology 54 Elsevier, 2019, pp. 112–121
  • [28] Louis K Scheffer et al. “A connectome and analysis of the adult Drosophila central brain” In Elife 9 eLife Sciences Publications Limited, 2020, pp. e57443
  • [29] Annalisa Scimemi “The Role of Astrocytes in Neurotransmitter Uptake and Brain Metabolism” In Computational Glioscience One New York Plaza, Suite 4600, New York, USA: Springer, 2019, pp. 309–328
  • [30] Alexey Semyanov “Spatiotemporal pattern of calcium activity in astrocytic network” In Cell Calcium 78 Elsevier, 2019, pp. 15–25
  • [31] Ann E Sizemore et al. “Cliques and cavities in the human connectome” In Journal of computational neuroscience 44.1 Springer, 2018, pp. 115–145
  • [32] Ann E Sizemore, Jennifer E Phillips-Cremins, Robert Ghrist and Danielle S Bassett “The importance of the whole: topological data analysis for the network neuroscientist” In Network Neuroscience 3.3 MIT Press, 2019, pp. 656–673
  • [33] Larry W Swanson and Jeff W Lichtman “From Cajal to connectome and beyond” In Annual Review of Neuroscience 39 Annual Reviews, 2016, pp. 197–216
  • [34] Jun Tang, Jin-Ming Luo and Jun Ma “Information transmission in a neuron-astrocyte coupled model” In PLoS One 8.11 Public Library of Science, 2013, pp. e80324
  • [35] Emma K Towlson et al. “The rich club of the C. elegans neuronal connectome” In Journal of Neuroscience 33.15 Soc Neuroscience, 2013, pp. 6380–6387
  • [36] Alexei Verkhratsky and Alexandr Chvátal “NMDA receptors in astrocytes” In Neurochemical Research 45.1 Springer, 2020, pp. 122–133
  • [37] Alexei Verkhratsky, José J Rodríguez and Vladimir Parpura “Calcium signalling in astroglia” In Molecular and cellular endocrinology 353.1-2 Elsevier, 2012, pp. 45–56
  • [38] Joshua T Vogelstein et al. “A community-developed open-source computational ecosystem for big neuro data” In Nature methods 15.11 Nature Publishing Group, 2018, pp. 846–847
  • [39] John J Wade et al. “Bidirectional coupling between astrocytes and neurons mediates learning and dynamic coordination in the brain: a multiple modeling approach” In PloS one 6.12 Public Library of Science, 2011, pp. e29445
  • [40] Keimpe DB Wierda et al. “The soluble neurexin-1β\beta ectodomain causes calcium influx and augments dendritic outgrowth and synaptic transmission” In Scientific reports 10.1 Nature Publishing Group, 2020, pp. 1–15
  • [41] Meng Xu et al. “Computer assisted assembly of connectomes from electron micrographs: application to Caenorhabditis elegans In PLoS One 8.1 Public Library of Science San Francisco, USA, 2013, pp. e54050
  • [42] Eleftherios Zisis et al. “Digital reconstruction of the neuro-glia-vascular architecture” In Cerebral Cortex 31.12 Oxford University Press, 2021, pp. 5686–5703