Skip to main content
Proceedings. Mathematical, Physical, and Engineering Sciences logoLink to Proceedings. Mathematical, Physical, and Engineering Sciences
. 2022 Apr 27;478(2260):20210828. doi: 10.1098/rspa.2021.0828

A mathematical model of network elastoplasticity

Hiroki Kodama 1,2, Ken'ichi Yoshida 3,
PMCID: PMC9044121  PMID: 35510220

Abstract

We introduce a mathematical model, based on networks, for the elasticity and plasticity of materials. We define the tension tensor for a periodic graph in a Euclidean space, and we show that the tension tensor expresses elasticity under deformation. Plasticity is induced by local moves on a graph. The graph is described in terms of the weights of edges, and we discuss how these weights affect the plasticity.

Keywords: polymer networks, periodic weighted graphs, discrete harmonic maps

1. Introduction

The field of topological crystallography was initially introduced by Kotani & Sunada [15] as a part of discrete geometric analysis. One of the main objects of their study is a net, that is, a periodic graph realized in RN. The energy of a net is defined as an analogue of the Dirichlet energy of a Riemannian manifold. In other words, one can say that the energy of a net is the total potential energy of springs, viewing edges as linear springs with rest lengths equal to zero. Harmonic and standard nets are defined as energy-minimizing nets under certain conditions, and they are regarded as equilibrium states. Nets have been used as models of crystals.

In this paper, we suggest that the energy of a net induces a model of hyperelastic materials. Here, hyperelasticity is the property from which stress under deformation is derived using an energy density function. To describe the deformation of a net, we introduce the notion of a tension tensor, which is regarded as multivariate energy. Further, a standard net is characterized by the tension tensor. We show that the Cauchy stress tensor is also expressed by the tension tensor. Furthermore, if the graph structure is preserved, the elasticity at the macro-scale is also determined by the tension tensor; otherwise, a departure from elasticity, known as plasticity, occurs. To describe the manner in which the graph structure changes, we consider two types of local moves: contraction and splitting. We define a condition for a local move and introduce two models of deformation concerning plasticity. This enables us to draw the stress–strain curve.

Our model is motivated by the structure of thermoplastic elastomers (TPEs). A TPE is a polymeric material with rubber elasticity and is remoldable at high temperatures. A typical TPE consists of ABA triblock copolymers, in which monomers of types A and B are arranged in a sequence such as AABBAA. ABA triblock copolymers of a certain type form two domains consisting of monomers A and B. This structure is called microphase separation. We consider a structure such that each component consisting of monomer A is a ball, as shown in figure 1. This is called a spherical structure. The domains consisting of monomers A and B are called hard and soft domains, respectively. The theoretical and numerical treatments of block copolymers are explained in the book by Fredrickson [6], to which we refer the reader for further information.

Figure 1.

Figure 1.

A spherical structure formed from ABA triblock copolymers.

In our model, the hard and soft domains correspond to the vertices and edges, respectively. More precisely, a hard domain is a vertex, and a polymer chain in the soft domain is an edge. The endpoints of an edge are the (possibly single) hard domains that contain monomer A of the copolymer. The obtained graph may have loops and multiple edges.

The network structure of polymers induces rubber elasticity. The random motion of polymer chains in the soft domain gives rise to entropic forces. A hard domain functions as a cross-link. In our approximation, we ignore the maximal length of the chain and the interaction between chains. If a chain moves randomly, the tension on the chain is proportional to the distance between the endpoints. This setting is consistent with the definition of the energy of a net. Suppose that the polymers can move freely while preserving the network structure. Additionally, we obtain a harmonic net in equilibrium. Since harmonicity is preserved by affine deformation, the affine assumption in the classical theory of rubber elasticity holds. Additional details of rubber elasticity are provided in the book by Treloar [7].

The hard domains of a TPE are less robust than the cross-links of vulcanized rubber because each hard domain is aggregated by intermolecular forces. The network structure of a TPE may change under deformation, as observed in simulation [8,9] and by conducting experiments [10]. For example, a hard domain may split, as shown in figure 2. Further, a loop may become a non-loop edge between the new domains. Conversely, two hard domains may contract. These moves cause plasticity. Although other moves may occur, we consider only contractions and splittings.

Figure 2.

Figure 2.

Splitting of a hard domain. In this example, one loop becomes a non-loop edge between the new domains.

In §2, we give preliminary definitions of nets. The graphs we use are weighted, and their weights can be regarded as the number of edges. Different types of polymers may contribute different weights.

In §3, we introduce the tension tensor. This is visualized by an ellipsoid.

In §4, we consider the elasticity of nets under deformation. Based on a physical argument, the Cauchy stress tensor is derived from the tension tensor. We also consider the stress under uniaxial extension. Young’s modulus of a standard net is determined by the energy per unit volume.

In §5, we define local moves. The contraction of two vertices is a natural operation. A splitting of a vertex is an inverse operation of contraction. The sum of weights is preserved in our model. The conditions for the occurrence of these moves are provided by the realization of a graph. The local tension tensor is used to determine whether a vertex splits or not. We suspend the physical validity of the conditions.

In §6, we introduce two models we call fast and slow deformations. Although these models reflect the dependence on the speed of deformation, we consider only the two extreme cases. Subsequently, we obtain the stress–strain curve, which is merely piecewise continuous. We suppose that the local moves under deformation finish in finitely many times. For example, after a vertex splits, the inverse contraction should not occur immediately. We show only a sufficient condition to avoid such repetitions.

Section 7 is devoted to mathematical results. In there, we consider the extent to which the weight of an edge affects the harmonic realization. For the sake of theoretical consideration, we allow weights to be non-negative real numbers and continuously deformed. When the weight of an edge in a harmonic net becomes large, the limit of nets is obtained by the contraction of this edge. In theorem 7.2, we show a mathematical result on the relation between the edge length and the difference of the tension tensor. Moreover, in theorem 7.6 we obtain a lower bound for the edge length.

In §8, we define the number called the energy loss ratio to measure the plasticity of nets. This provides an estimate of the permanent strain for uniaxial tension. We observe simple examples, which suggest the following:

  • (i)

    a material has lower plasticity if the proportion of loops is large;

  • (ii)

    a material with lower plasticity is obtained by blending two materials.

Continuous deformation of weights highlights these tendencies.

In §9, we give examples of deformation. We use a periodic graph obtained from the hexagonal lattice. The nets obtained by deformation depend on the stretching direction.

2. Definitions

Based on the formulation in [3,5], we prepare some notions in topological crystallography. Let X=(V,E,w) be an (abstract) weighted graph, which is defined by the vertex set V and the edge set E with maps o,t:EV, and ι:EE such that ι2=id, ι(e)e, and o(ι(e))=t(e) for any eE. The maps o and t associate the origin and the terminal of an edge, respectively. The map ι reverses the orientation of an edge. We allow a loop (edge e such that o(e)=t(e)) and a multi-edge (edges with common terminal points). The weight function w:ER0 satisfies w(ι(e))=w(e) for eE. We regard the weight of an edge as the number of edges. Hence, we may replace an edge e0 with the union of edges e1 and e2 if o(e0)=o(e1)=o(e2),t(e0)=t(e1)=t(e2), and w(e0)=w(e1)+w(e2). The weight function is often omitted in the notation. The degree of a vertex vV is defined by deg(v)=o(e)=vw(e). Note that the weight of a loop contributes twice to the degree of its endpoint.

A graph X=(V,E) is a finite graph if V and E are finite sets. Otherwise, X is an infinite graph. We can naturally identify X with a one-dimensional complex. Note that two elements e and ι(e) in E correspond to a single 1-cell in the complex. We may reduce the complex by removing the zero-weight edges. If this reduced complex is connected, we say that the graph is connected.

We consider an infinite connected graph X. For N1, suppose that L=ZN acts on X as (weight-preserving) automorphisms of the graph, the quotient map ω:XX/L=(V/L,E/L) is a covering, and X/L is a finite graph. Then, we say that X is a periodic graph, and L is a period lattice for X. A map Φ:VRN is called a periodic realization of X in RN if there exists an injective homomorphism ρ:LRN as Z-modules satisfying that

  • (i)

    Φ(γv)=Φ(v)+ρ(γ) for any vV and γL, and

  • (ii)

    ρ(L) is a lattice subgroup of RN.

Condition (i) means that Φ is L-equivariant. We call ρ and ρ(L), respectively, the period homomorphism and the period lattice for Φ.

Definition 2.1. —

The pair (X,Φ) is called as a net in RN if Φ is a periodic realization of a periodic graph X in RN.

A periodic realization Φ maps an edge eE to a vector

vΦ(e)=Φ(t(e))Φ(o(e)),

in RN. Since vΦ(γe)=vΦ(e) for γL, we obtain vΦ:E/LRN by vΦ(ω(e))=vΦ(e). We often write Φ instead of vΦ by abuse of notation. If e is a loop, then Φ(e)=0.

We define the energy of a net and consider energy-minimizing realizations. Note that our definition of the energy is slightly different from that in [5, §7.4], where the energy normalized by the volume is defined.

Definition 2.2. —

The energy (per period) of a net (X,Φ) is defined as follows:

E(X,Φ)=12eE/Lw(e)||Φ(e)||2.

In other words, when we regard edges as springs, the energy is two times the total potential energy of linear springs with rest length equal to zero and elasticity constant given by the edge weight. Note that we count the segment between points P and Q twice in the summation, as edges from P to Q and from Q to P.

Definition 2.3. —

A periodic realization Φ of X is called harmonic if the energy E(X,Φ) is minimal among the periodic realizations of X with the fixed period homomorphism ρ. Then, we call (X,Φ) a harmonic net.

Definition 2.4. —

A periodic realization Φ of X is called standard if the energy E(X,Φ) is minimal among the periodic realizations of X with the fixed covolume vol(RN/ρ(L)). Then, we call (X,Φ) a standard net.

Remark that when vol(RN/ρ(L))=vol(RN/ρ~(L)), there exists a volume-preserving linear transformation ASL(N,R) satisfying Aρ=ρ~. Therefore, a periodic realization is standard if its energy is minimal among its volume-preserving linear transformations.

Clearly, a standard realization is harmonic. A harmonic realization is characterized by a local condition. We characterize a standard realization in §3.

Theorem 2.5. ([5] theorem 7.3) —

A periodic realization Φ is harmonic if and only if o(e)=vw(e)Φ(e)=0 for any vV.

From this theorem, it directly follows that a linear transformation of a harmonic representation is also harmonic.

Corollary 2.6. —

Suppose that Φ is a periodic realization and AGL(N,R) is a linear transformation Then the composition AΦ is a harmonic realization if and only if Φ is harmonic.

Remark 2.7. —

One might think that in definition 2.2, the rest lengths of the springs should be positive, not zero. However, we have assumed the rest lengths to be zero for two reasons. The first reason is due to statistical mechanics. A chain in a TPE is not taut like a helical spring, but it fills space randomly. Therefore, the tension on the chain is proportional to the distance between the endpoints. The second reason is a mathematical one. We will transform harmonic nets by continuous linear transformations in the following sections. However, in order for the result of any linear transformation to be harmonic again, the natural length of the spring must be zero (see corollary 2.6).

Remark 2.8. —

In this paper, we use the term ‘net’ as a periodically realized network in RN. In the book of Wells [11], who initiated a systematic study of crystal structures as networks, a connected simple periodic graph with straight edges in a Euclidean space was called a net, and we follow this convention. Note that in the terminology of [5], a net is called a topological crystal.

3. Tension tensor

In our mathematical model, a net represents the structure of TPE chains. Consider the tension caused by the structure. Indeed, a stretched TPE must have tension in the direction in which the structure is stretched. For example, the net in figure 3 has a symmetric shape. Thus, it has no tension in any direction. By contrast, the net in figure 4 seems to be stretched from the top right to the bottom left. However, what can we say about a more complicated net such as that in figure 5?

Figure 3.

Figure 3.

A symmetric net with tension ellipse. (Online version in colour.)

Figure 4.

Figure 4.

A stretched net with tension ellipse. (Online version in colour.)

Figure 5.

Figure 5.

A complicated net with tension ellipse. (Online version in colour.)

To answer this question, we introduce a matrix named a tension tensor. Essentially, the tension tensor denotes the energy of a net with information of the direction along which the net is stretched, or its ‘directed energy’. We will observe that the tension tensor can be visualized as an ellipse (or ellipsoid), such as the ellipses in figures 35.

Through computer simulation for the deformation of two-dimensional nets, we make the following observations:

  • (i)

    when we stretch a net and the graph structure does not experience any change, the ellipse of the tension tensor also stretches in the same direction; and

  • (ii)

    when the graph structure changes, the ellipse of the tension tensor becomes round.

The former immediately follows from the definition. The latter can be verified by theorem 7.2.

(a) . Definition of the tension tensor

Definition 3.1. —

For a net (X,Φ), we define the local and global tension tensors as follows: For a vertex v of X or X/L, the local tension tensor is defined by

T(v)=o(e)=vw(e)Φ(e)2,

where

(x1xN)2=(x1xN)(x1xN)=(x1xN)(x1,,xN)=(x12x1xNxNx1xN2).

The global tension tensor (per period) is defined by

T(X,Φ)=12vV/LT(v).

Proposition 3.2. —

It holds that tr(T(X,Φ))=E(X,Φ).

Proof. —

tr(T(X,Φ))=12vV/Lo(e)=vw(e)tr(Φ(e)2)=12eE/Lw(e)||Φ(e)||2=E(X,Φ).

Note that two elements e,ι(e)E/L are distinguished.

The following characterization of a standard realization follows from ([5], Theorem 7.5).

Theorem 3.3. —

A periodic realization Φ is standard if and only if it is harmonic and the global tension tensor T(X,Φ) is a constant multiple of the identity matrix.

Consequently, a standard realization is unique up to similar transformations. The existence and explicit constructions of a standard realization were also shown previously [3,5].

We remark that the global tension tensor T(X,Φ) per period depends on the choice of period. Suppose that L2 is a finite index sublattice of L1=L, and Ti is the tension tensor with respect to the lattice Li. Then

T2(X,Φ)=[L1:L2]T1(X,Φ).

To avoid this ambiguity, we can define the tension tensor per weight by

Tw(X,Φ)=T(X,Φ)12eE/Lw(e).

However, in most parts of this paper, we assume that the covolumes of period lattices are constant, and we use the tension tensor per period without the ambiguity.

(b) . Linear action and visualization

Let AGL(N,R). The matrix A acts on a net (X,Φ) by

A(X,Φ)=(X,AΦ).

Since xx=xxT for xRN, it follows that T(A(X,Φ))=AT(X,Φ)AT, where xT and AT are the transposes of x and A, respectively. In particular, if A is a symmetric matrix, then T(A(X,Φ))=AT(X,Φ)A.

To visualize the tension tensor, we define an ellipsoid by

Ell(X,Φ)={xRNxTTw(X,Φ)1x=1}.

We remark that we use the tension tensor per weight here to avoid ambiguity. It is easy to check that Ell(A(X,Φ))=AEll(X,Φ).

4. Stress

In this section, we consider the stress experienced by a net by using the tension tensor. Fix a periodic graph X. Let Φ be a harmonic realization of X. We regard the energy E=E(X,Φ) as physical energy. This is interpreted as the Helmholtz free energy for entropic elasticity. With a three-dimensional object in mind, we give an obvious generalization to the N-dimensional version. (See [12] for the classical theory on continuum mechanics.)

As a result, the stress satisfies the neo-Hookean model, which is the simplest one among the hyperelastic materials. This also coincides with the consequence of the classical theory on rubber elasticity by Kuhn (see [7], ch. 4). Note that his setting is not identical to ours. Although polymers are normally distributed in his theory, the net in our setting is not isotropic. Nonetheless, a standard net has isotropy at the macro-scale.

(a) . Stress tensor

By compositing rotational isometry, we may assume that the tension tensor per period is a diagonal matrix T=T(X,Φ)=diag(τ1,,τN). Then, the energy per period is E=trT=iτi. Let V=vol(RN/ρ(L)) denote the volume per period.

We apply a physical argument to define stress for nets. Let us consider a macro-scale object of which the shape is an orthotope (N-cuboid) of edge length Li in each ith direction for 1iN. Suppose that this object consists of a net at the micro-scale. We apply the affine deformation assumption ([7], ch. 4) (or the Cauchy–Born rule [13]). In other words, if the macro-scale object undergoes an affine deformation, the net at the micro-scale undergoes the same affine deformation. Then, the total energy is equal to (iLi/V)E=(iLi/V)iτi. Suppose that external force Fi extends outward in each ith direction, and the object remains in equilibrium. Then, the stress in the ith direction is σi=LiFi/jLj. Consider infinitesimal deformation of the object. For a short while, we allow the volume V to vary but let iLi/V be constant. Let ΔLi denote the displacement in the ith direction. The strain in the ith direction is ϵi=ΔLi/Li. Then, the work is iFiΔLi=jLjiσiϵi, which is equal to the difference of energies (jLj/V)ΔE. Hence, iσiϵi=ΔE/V. The difference of the tension tensor is given by

ΔT=diag(τ1(1+ϵ1)2,,τN(1+ϵN)2)T=diag(2τ1ϵ1,,2τNϵN),

modulo the order more than one. Hence, ΔE=tr(ΔT)=i2τiϵi. Therefore,

iσiϵi=i2τiVϵi.

If we can vary ϵi freely, we obtain σi=2τi/V. Thus, we define the Cauchy stress tensor for a net as Σ=(2/V)T, which is valid in general coordinates.

Furthermore, we suppose that the deformation preserves the volume. In other words, iLi and V are constant. Since i(Li+ΔLi)=iLi, we have iϵi=0. If we vary ϵi under this condition, the equation iσiϵi=i(2τi/V)ϵi implies that σi=2τi/Vc for some constant c. Indeed, uniform pressure does not change the shape under the constraint of volume. Thus, the traceless part of the Cauchy stress tensor Σ(tr(Σ)/N)I=(2/V)T(2E/NV)I is regarded as the volume-preserving part. This is called the deviatoric stress tensor.

If the external forces Fi are equal to zero, then 2τi/V=c. Hence, the deviatoric stress tensor is zero. Since T=(cV/2)I, theorem 3.3 implies that Φ is a standard realization.

A material is hyperelastic (or Green elastic) if the stress under deformation is determined by a strain energy density function. In our setting, consider the affine deformation of a standard net (X,Φ) by the diagonal matrix

A=diag(λ1,,λN)SL(N,R).

The strain energy density function is given by

1V(E(A(X,Φ))E(X,Φ))=E(X,Φ)NV(iλi2N).

A material with such a strain energy density function is called incompressible neo-Hookean.

(b) . Uniaxial extension

Consider a harmonic net (X,Φ). For the sake of the argument in §8, first let Φ be not necessarily standard. We write (τij)1i,jN=T(X,Φ), which is not necessarily diagonal, in contrast to the previous subsection. For λ>0, the diagonal matrix

A(λ)=diag(λ,λ1/(N1),,λ1/(N1))SL(N,R),

induces a uniaxial extension with strain ϵ=λ1. The volume V per period is constant under deformation. Consider the tension tensor T(λ)=T(A(λ)(X,Φ))=A(λ)T(X,Φ)A(λ) after deformation. A stress tensor in the volume-preserving setting is given by (σij)1i,jN=(2/V)T(λ)cI for some c. By considering the nature of uniaxial extension, we suppose that σ22++σNN=0. However, it does not hold that σ22==σNN=0 in general. Then,

c=2(N1)V(τ22++τNN)λ2/(N1).

The true stress under this uniaxial extension is defined by

σtrue=σ11=2Vτ11λ2c=2V(τ11λ21N1(τ22++τNN)λ2/(N1)).

The engineering stress (or nominal stress) is measured using the cross-sectional area before deformation, and it is defined by σeng=σtrue/λ.

Proposition 4.1. —

Let E(λ)=E(A(λ)(X,Φ)). Then

σtrue=λVdE(λ)dλ,σeng=1VdE(λ)dλ.

Proof. —

Since E(λ)=tr(T(λ))=τ11λ2+(τ22++τNN)λ2/(N1), we have

dE(λ)dλ=2(τ11λ1N1(τ22++τNN)λ12/(N1)).

The permanent strain is the number ϵ0 satisfying σeng(1+ϵ0)=0. The following equality is clear from the definition of σeng.

Proposition 4.2. —

It holds that

ϵ0=(τ22++τNN(N1)τ11)(N1)/2N1.

Consider the case in which Φ is standard. Then, τ11==τNN=E(X,Φ)/N. Moreover, we have σ22==σNN=0, which is natural for uniaxial extension. The true stress is given by

σtrue=σ11=2E(X,Φ)NV(λ2λ2/(N1)).

The engineering stress is given by

σeng=σtrueλ=2E(X,Φ)NV(λλ12/(N1)).

Then, Young’s modulus for a standard net (X,Φ) is defined by

E=dσtruedλ|λ=1=4E(X,Φ)(N1)V.

5. Local moves

We introduce three local moves for nets: contraction and splitting. A local move for a graph is an operation to obtain a new graph by replacing some vertices and edges. When we say that we replace an edge e with e, we simultaneously replace ι(e) with ι(e) in our notation. For a periodic graph, a local move is regarded as an equivariant operation preserving the period. Even though local moves are defined as operations for abstract graphs, the conditions under which they occur are given by realizations of nets.

(a) . Contraction of two vertices

Let X=(V,E) be an (abstract) graph, and let v0, v1V. We construct a new graph X=(V,E) as follows: We define V=(V{v0,v1})v and E=E. Let π:VV denote the projection such that π(v0)=π(v1)=v and the restriction π|V{v0,v1} is the identity map. We define the endpoint maps o=πo,t=πt:EV. Suppose that the weight function on E is identical to that on E. We call this operation the contraction of v0 and v1 to v. This contraction causes the edges e0,e1 and e01 to change into the loop e on v, where ei is the loop on vi, and e01 is the edge between v0 and v1 (figure 6). Then, the sum of weights is preserved.

Figure 6.

Figure 6.

Contraction of two vertices.

For a periodic graph X with period L, we define a contraction as an equivariant operation. In other words, we apply the contraction of γv0 and γv1 for each γL. In this case, it is necessary that v1γv0 for any γL. As a result, we obtain a new periodic graph X.

To introduce deformation in §6, we define a condition for contraction using a realization Φ of the graph X. Fix a constant δ>0. If ||Φ(v1)Φ(v0)||δ, then we suppose that the vertices v0 and v1 contract.

Note that the weight w(e01) may be zero. Even in this case, the vertices v0 and v1 contract if their distance is sufficiently small.

(b) . Splitting of a vertex

A splitting of a vertex is an inverse operation of contraction. This is not determined only by the vertex. When a vertex v splits, the loop on v changes the loops on v0 and v1 and the edge between them. Although the sum of weights is preserved, the choice of the three weights is not unique. We also need to assign an endpoint v0 or v1 for each new edge corresponding to an edge originating from v. Note that there may not necessarily exist loops on the vertex v. If there exist no loops on v, there exist no edges between the new vertices v0 and v1.

For a periodic graph X, we define a splitting as an equivariant operation to obtain a new periodic graph X. We remark that vertices v and γv for γL may be adjacent. Even in this case, we can still define a splitting as such. By ignoring the period, we apply successive splittings of v and γv. Note that the sequence of splittings in any order yields the same result.

We define a condition for splitting using a realization Φ of the graph X:

  • (i)

    when a vertex v splits;

  • (ii)

    the way in which the edges originating from v are divided into two classes; and

  • (iii)

    the way in which the weights are assigned.

Fix constants Kd>0 for d>0. The value Kd is regarded as the firmness of a vertex with degree d. Recall that Φ:ERd is the map induced by Φ, and T(v)=o(e)=vw(e)Φ(e)2 is the local tension tensor around v. Let λmax denote the maximal eigenvalue of T(v). Take an eigenvector u associated with λmax. We divide the edges originating from v into two classes {e0,j}1jm and {e1,j}1jn so that uΦ(e0,j)0 and uΦ(e1,j)0. If λmaxKdeg(v), then we suppose that the vertex v splits. This may be regarded as the maximal principal stress criterion. The new edge corresponding to ei,j originates from vi. Roughly speaking, the splitting occurs in the stretched direction.

To obtain the unique division into two classes {e0,j} and {e1,j}, we need the following condition of genericity:

  • (i)

    the eigenspace associated with λmax is the one-dimensional space span(u), and

  • (ii)

    there are no edges ei,j such that uΦ(ei,j)=0.

Because it is difficult to decide how the weights are assigned, we use an ad hoc setting: suppose that e0 and e1 are, respectively, the loops on v0 and v1. Moreover, suppose that e01 is the edge between v0 and v1. Fix the probabilities p0, p1 and p01 that the loop e on v changes into the new edges e0, e1 and e01. In other words, w(e0)=p0w(e), w(e1)=p1w(e), w(e01)=p01w(e) and p0+p1+p01=1. Although the choice of p0, p1 and p01 may be arbitrary, it is reasonable to set p0=p1=1/4 and p01=1/2. The reason is that the above choice holds if each endpoint of a new edge is v0 with a probability of 1/2.

For a realization Φ of a graph X, suppose that a graph X is obtained by splitting a vertex v into v0 and v1. Then, we define the immediate realization Φ(i) of X (or Φ by abuse of notation) as follows: Φ(i)(v0)=Φ(i)(v1)=Φ(v), and Φ(i)(u)=Φ(u) for any other vertex u. If Φ is a periodic realization, then Φ(i) is equivariantly defined as a periodic realization. Using this, we can show that splitting decreases the energy.

Proposition 5.1. —

Suppose that X is a graph obtained from X as a result of splitting under the above condition. Let Φ be a harmonic realization of X with the same period as Φ. Then, E(X,Φ)<E(X,Φ).

Proof. —

Clearly, E(X,Φ(i))=E(X,Φ). The condition of splitting and theorem 2.5 imply that Φ(i) is not harmonic. Hence, E(X,Φ)<E(X,Φ(i)) by definition 2.3.

Remark 5.2. —

In the above two subsections, the graph X obtained by the local move is well defined. However, the realization Φ of X should be defined using harmonicity, so there remains ambiguity of parallel translation.

In this paper, it does not matter because we only consider the shape of the realized graph and its energy. However, if one wants to discuss such as the displacements of nodes before and after a local move, this ambiguity should be removed. One idea is to assume that the centre of mass of the periodic cell is fixed.

6. Models of deformation

We introduce two models: fast and slow deformation. The difference of these two models reflects the strain rate sensitivity, that is, the dependency of stress on the speed of deformation. A harmonic realization of a periodic graph is regarded as an equilibrium state. Let (X0,Φ0) be a standard net with period homomorphism ρ0 as an initial condition. This is regarded as a state without external force, as explained in §4. In this section, we express the period homomorphisms explicitly. Suppose that the initial net (X0,Φ0,ρ0) does not satisfy any condition of a contraction or splitting. Deformation is obtained by linear transformations with constant volume. We apply contractions and splittings that satisfy the conditions in §5 for harmonic nets in deformation. Because the process is not deterministic in general, it is necessary to choose one that satisfies the conditions. We leave the stochastic formulation for future work.

(a) . Fast deformation

Let ASL(N,R). Fix the period homomorphism Aρ0. We apply local moves for the harmonic net A(X0,Φ0,ρ0)=(X0,AΦ0,Aρ0). First, we apply the splittings. Then, we obtain a new harmonic net. If the conditions of other local moves hold, we continue to apply the splittings. Second, we apply the contractions. However, more than two vertices may contract to a point. In general, we need to choose the contracting vertices so that the contraction does not violate the period.

We continue this procedure by supposing that these procedures finish with finitely many local moves. In the end, we obtain a harmonic net (X1,Φ1,Aρ0). We call this process fast deformation.

(b) . Slow deformation

Slow deformation is a limit of sequences of fast deformation. For a continuous family of linear transformations, take approximations by discrete families of small ones. They induce sequences of fast deformation. We obtain slow deformation by the limit as the approximations get arbitrarily fine.

Equivalently and more precisely, slow deformation is defined as follows. Suppose that AtSL(N,R) for 0t1 is a continuous family of linear transformations such that A0=id. Let ρt=Atρ0. We apply local moves while increasing t from zero to one. Let t1 be the minimal t such that the condition of a local move holds for the harmonic net (X0,AtΦ0,ρt). We obtain a graph X0 by the local move. Consider a harmonic net (X0,Φ0,ρt1). Note that another local move may occur for (X0,Φ0,ρt1). Then, we continue to apply local moves. Subsequently, we obtain a harmonic net (Xt1,Φt1,ρt1).

After the exhaustion of local moves for t1, we increase t. Let t2 be the minimal t more than t1 such that the condition of a local move holds for the harmonic net (Xt1,AtAt11Φt1,ρt1). Using the same argument as above, we obtain a harmonic net (Xt2,Φt2,ρt2).

We continue this procedure by supposing that these procedures finish with finitely many local moves. Ultimately, we obtain a harmonic net (X1,Φ1,ρ1). We call this process slow deformation.

A condition of genericity is given as follows:

  • (i)

    two local moves do not occur simultaneously, and

  • (ii)

    two vertices equivalent by the period do not contract.

If the net is highly symmetric, the genericity is difficult to hold. For genericity, we arbitrarily choose a single local move at a time, and we ignore any contraction that violates the period.

We define the stress–strain curve for a uniaxial extension. Let

A(λ)=diag(λ,λ1/(N1),,λ1/(N1))SL(N,R).

The slow deformation by At=A(λt) for 0t1 induces a net (X1,Φ1). The energy E(λ)=E(X1,Φ1) is a right-continuous function of λ. We can plot the stress–strain curve as a graph of the engineering stress σeng as a function of the strain ϵ=λ1, where σeng=(1/V)(dE(λ)/dλ) by proposition 4.1.

(c) . Compatibility of splitting and contraction

We suppose that the above procedures in deformation finish with finitely many local moves. In general, splittings and contractions may cause an infinite sequence of local moves. We give only a partial result for this problem.

Let (X,Φ) be a harmonic net. Let X be a periodic graph obtained from X by splitting a vertex v into v0 and v1 with respect to the condition given in §5. The maximal eigenvalue of the local tension tensor T(v) is equal to Kdeg(v). Let e denote the new non-loop edge in X. Let w be the weight of e, which does not exceed the weight of the loop on v. We consider a harmonic realization Φ with the same period lattice as Φ. If ||Φ(e)||=||Φ(v1)Φ(v0)||δ, then splittings and contractions continue alternately. We show that such repetition does not occur if δ is sufficiently small.

Theorem 6.1. —

Suppose that the weights are non-negative integers. Then,

||Φ(e)||2Kdeg(v)deg(v).

Proof. —

Let Φ(a) be an auxiliary periodic realization of X such that Φ(a)(v0)=Φ(v), Φ(a)(u)=Φ(u) for any vertex uγv1 (γL), and Φ(a) is harmonic around v1. Then, ||Φ(e)||||Φ(a)(e)||, which we show in theorem 7.6. Hence, it is sufficient to show that

||Φ(a)(e)||1deg(v1)Kdeg(v)2.

Indeed, since deg(v)=deg(v0)+deg(v1), we may assume that deg(v)2deg(v1) by interchanging v0 and v1 if necessary.

Let ei1,,eini for i=0,1 denote the non-loop edges of X originating from vi other than e, and let vi1,,vini denote their terminals. Note that we may ignore the edges between vi and γvi for γL. Let wij be the weight of eij. The same symbol is used for the corresponding edges and vertices of X. We write Φ(vij)=(xij1,,xijN). We may assume that Φ(v)=Φ(a)(v0)=0 and the splitting occurs in the direction of the first coordinate; that is, the vector (1,0,,0) is an eigenvector associated with the maximal eigenvalue of the local tension tensor around v. Then, x0j10, x1j10, i,jwijxij1=0, and i,jwij(xij1)2=Kdeg(v). Since Φ(a) is harmonic around v1, we have

wΦ(a)(v1)+j=1n1w1j(Φ(v1j)Φ(a)(v1))=0.

Hence,

||Φ(a)(e)||=||Φ(a)(v1)||=||j=1n1w1jΦ(v1j)w+j=1n1w1j||j=1n1w1jx1j1deg(v1).

Moreover, we obtain j=1n1w1jx1j1Kdeg(v)/2 by lemma 6.2 and the assumption that wijZ0.

Lemma 6.2. —

Let x01,,x0n00 and x11,,x1n10. Suppose that z=j=1n0x0j=j=1n1x1j and K=i,j(xij)2. Then, zK/2.

Proof. —

We find the maximum K for a fixed z>0. If x+y=a is fixed for x,y0, then the maximum a2 of x2+y2 is attained when x=0 or y=0. Hence, the maximum of K is attained when x0j0=z and x1j1=z for some j0 and j1, and xij=0 for the other j. Therefore, K2z2.

7. Variation of weights

In this section, we consider the extent to which the harmonic realizations and their energies depend on the weights, which vary in non-negative real numbers. We describe a contraction as the limit by increasing the weight of an edge. Note that we do not suppose that the sum of weights is preserved, which differs from the assumption in §5.

Let X=(V,E) be a periodic graph with period L. We take representatives v0,v1,,vnV of the set V/L. Let eijγ denote the edge from vi to γvj for γL. Then, {eijγ} are representatives of the set E/L. Let wijγ denote the weight of eijγ. Then, wij,γ=wjiγ.

Fix a period homomorphism ρ:LRN. Suppose that n1. Let X^ be a periodic graph obtained from X by the contraction of v0 and v1 to a vertex v^. Suppose that Φ(h) and Φ^(h) are harmonic realizations of X and X^, respectively. We may assume that Φ(h)(v0)=Φ^(h)(v^)=0. We change the harmonic realizations Φ(h) by varying w010 while fixing the other weights on X. Here, we regard wijγ=wji,γ as a single variable. Suppose that X is connected, that is, the union of its edges with positive weights is connected. Since the harmonic realization Φ(h) is given by the unique solution of a system of linear equations, it depends continuously on w010. After we show that Φ^(h) can be regarded as limw010Φ(h), we give explicit presentations.

Lemma 7.1. —

The realizations Φ(h) converge to Φ^(h) as w010. In other words, Φ(h)(vi) converge to Φ^(h)(vi) for each 2in, and Φ(h)(v1) converge to Φ^(h)(v^)=0. In particular, limw010Φ(h)(e010)=0. Consequently,

limw010T(X,Φ(h))=T(X^,Φ^(h))andlimw010E(X,Φ(h))=E(X^,Φ^(h)).

Proof. —

Consider a (not necessarily harmonic) periodic realization Φ of X. We write Φ(vi)=(xi1,,xiN)RN and ρ(γ)=(ργ1,,ργN)RN for γL. Then, ργk=ργk. By theorem 2.5, the realization Φ is harmonic if and only if

j=0nγLwijγ(xik+xjk+ργk)=0, 7.1

for any 0in and 1kN. Let xik=ξik be a solution of this system of equations, which is unique up to translations. We may assume that ξ0k=0. Then, Φ(h)(vi)=(ξi1,,ξiN). We write wij=γwijγ, bij=wij for ij, bii=jiwij, and cik=j,γwijγργk. Using the matrix B=(bij), the system of linear equations is written as B(x0k,,xnk)T=(c0k,,cnk)T. Since ξ0k=0 and (ξik)1in are unique, we have B00(ξ1k,,ξnk)T=(c1k,,cnk)T, where B00=(bij)1i,jn is a minor of B. Then, B00 is invertible. Cramer’s rule implies that ξik=detCik/detB00, where Cik is the matrix obtained by replacing the ith column of B00 with (c1k,,cnk)T. In this presentation of detCik/detB00, only b11=j1,γw1jγ contains w010=w100. (Note that ρ0k=0.) Hence, detB00 and detCik are linear functions of w010. Moreover, detC1k is constant for w010.

Since Φ(h) is harmonic, we obtain E(X,Φ(h))E(X^,Φ^(h)) by regarding Φ^(h) as a realization of X. Moreover, we have w010||Φ(h)(e010)||2E(X,Φ(h)). Hence, limw010Φ(h)(e010)=0. In other words, limw010ξ1k=0. If detB00 is constant for w010, then ξ1k=detC1k/detB00 is also constant. Hence, ξ1k=0. Then, Φ(h)=Φ^(h), and the assertion holds trivially.

Suppose that detB00 is not constant for w010. Then, ξik=detCik/detB00 converges as w010. We define the realization Φ^ of X^ such that Φ^(v^)=0 and Φ^(vi)=limw010Φ(h)(vi) for 2in. Since E(X^,Φ^)=limw010E(X,Φ(h))E(X^,Φ^(h)) and Φ^(h) is harmonic, the realization Φ^ is also harmonic. The uniqueness of a harmonic realization implies that Φ^=Φ^(h). Therefore, limw010Φ(h)=Φ^(h).

Theorem 7.2. —

There are zRN and WR such that

Φ(h)(e010)=zw010+W

and

T(X^,Φ^(h))T(X,Φ(h))=z2w010+W,

where z and W do not depend on w010, and W does not depend on ρ. Consequently,

E(X^,Φ^(h))E(X,Φ(h))=||z||2w010+W=(w010+W)||Φ(h)(e010)||2.

Proof. —

We showed that ξ1k=detC1k/detB00 in the proof of lemma 7.1. The determinants detB00 and detC1k are, respectively, linear and constant as functions of w010=w100. If detB00 is constant for w010, then ξ1k=0, and we obtain z=0. We can take W arbitrarily.

Suppose that detB00 is not constant for w010. Then, we can write detB00=(w010+W)detB00,11, where B00,11=(bij)2i,jn, and W does not depend on w010 or ρ. Let zk=detC1k/detB00,11 and z=(z1,,zN). Then,

Φ(h)(e010)=(ξ11,,ξ1N)=zw010+W.

Moreover, z does not depend on w010.

We regard the tension tensor T(X,Φ) as a function of w=(wijγ) and x=(xik). Its (k,l)-entry is given by

Tkl(w,x)=12i,j=0nγLwijγ(xik+xjk+ργk)(xil+xjl+ργl).

Then,

Tklxαβ(w,x)=δβkj,γwαjγ(xαl+xjl+ργl)δβlj,γwαjγ(xαk+xjk+ργk).

Consider Tkl(w)=Tkl(w,ξ(w)) as a function of w, where ξ(w)=(ξik(w)). By the equality (7.1), we have Tklxαβ(w,ξ(w))=0. Hence,

Tklw010(w)=Tklw010(w,ξ(w))+α,βTklxαβ(w,ξ(w))ξαβw010(w)=Tklw010(w,ξ(w))=ξ1kξ1l=zkzl(w010+W)2,

where we regard w010=w100 as a single variable. Hence, Tkl=Czkzl/(w010+W) for some C independent of w010. Since limw010T(X,Φ(h))=T(X^,Φ^(h)) by lemma 7.1, we have

T(X^,Φ^(h))T(X,Φ(h))=(zkzlw010+W)1k,lN=z2w010+W.

Lemma 7.3. —

Let the vector z and the number W be as in theorem 7.2. Suppose that z0. Then,

0<W(j1γLw1jγ)w100.

Proof. —

Since

||z||2w010+W=E(X^,Φ^(h))E(X,Φ(h))<E(X^,Φ^(h))

for any w0100, we have W>0.

As in the proof of theorem 7.2, we have w010+W=detB00/detB00,11, where wij=γwijγ, bij=wij for ij, bii=jiwij, B00=(bij)1i,jn and B00,11=(bij)2i,jn. Then,

detB00detB00,11=b11(b12,,b1n)B00,111(b12,,b1n)T,

by lemma 7.4. Since B00,11 is positive definite by lemma 7.5 and detB00,110, so is B00,111. Therefore, w010+Wb11=j1γLw1jγ. Note that if n=1, we conventionally set detB00,11=1 and w010+W=b11.

Lemma 7.4. —

Let A=(aij)1i,jn be a symmetric matrix. Suppose that A11=(aij)2i,jn is invertible. Then

detAdetA11=a11(a12,,a1n)A111(a12,,a1n)T.

Proof. —

Using the adjugate matrix, we have

A111=(detA11)1((1)i+jdetA11,ji)2i,jn,

where A11,ij=(akl)k1,i,l1,j. The cofactor expansion implies that

detA=a11detA11+2i,jn(1)i+j+1ai1a1jdetA11,ij=a11detA11(a12,,a1n)(detA11)A111(a12,,a1n)T.

Lemma 7.5. —

Let A=(aij)1i,jn be a symmetric matrix. Suppose that aij0 for any ij and j=1naij0 for any i. Then, A is positive semi-definite.

Proof. —

The proof is by induction on n. The assertion is trivial for the case n=1. We have

A=diag(j=1na1j,,j=1nanj)+PTAP,

where

A=(a11a1,n10an1,1an1,n10000)diag(j=1na1j,,j=1nan1,j,0)

and

P=(101011001).

The matrix A is positive semi-definite by the assumption of induction for n1. Therefore, A is also positive semi-definite.

Theorem 7.6. —

Let Φ(a) be a realization of X such that Φ(a)(vi)=Φ^(h)(vi) for any i1 and Φ(a) is harmonic around v1. Then, ||Φ(h)(e010)||||Φ(a)(e010)||.

Proof. —

Since Φ(a) is harmonic around v1 and γLw11γρ(γ)=0, we have

j1γLw1jγ(Φ(a)(v1)+Φ^(h)(vj)+ρ(γ))=0.

Hence,

Φ(a)(v1)=j1γLw1jγ(Φ^(h)(vj)+ρ(γ))j1γLw1jγ.

Let

z(a)=j1γLw1jγ(Φ^(h)(vj)+ρ(γ))

and

W(a)=(j1γLw1jγ)w100.

Then, Φ(a)(e010)=Φ(a)(v1)=z(a)/(w010+W(a)). The difference of energies is given by

E(X^,Φ^(h))E(X,Φ(a))=j1γLw1jγ||Φ^(h)(vj)+ρ(γ)||2j1γLw1jγ||Φ(a)(v1)+Φ^(h)(vj)+ρ(γ)||2=j1γLw1jγ||Φ(a)(v1)||2+2j1γLw1jγ(Φ^(h)(vj)+ρ(γ))Φ(a)(v1)=(w010+W(a))||z(a)w010+W(a)||2+2z(a)z(a)w010+W(a)=||z(a)||2w010+W(a).

Since Φ(h) is harmonic, we have E(X,Φ(h))E(X,Φ(a)). Lemma 7.3 implies that WW(a). Therefore,

||Φ(h)(e010)||2=E(X^,Φ^(h))E(X,Φ(h))w010+WE(X^,Φ^(h))E(X,Φ(a))w010+W(a)=||Φ(a)(e010)||2

by theorem 7.2.

8. Plasticity

The plasticity of a material is its ability to undergo permanent deformation. For a fixed periodic graph, no external force is applied to a net if and only if its realization is standard, as shown in §4. Hence, if no local moves occur in the deformation of a net, then it returns to its initial state by unloading, and it is perfectly elastic. In general, however, local moves cause plasticity. To measure plasticity, we introduce the energy loss ratio of a net under deformation. This is defined by comparing the energy with that of a net in which local moves do not occur.

Definition 8.1. —

Let (X0,Φ0) be a harmonic net in RN. Let us consider fast deformation by A=A1SL(N,R) or slow deformation by AtSL(N,R) for 0t1. Suppose that we obtain a harmonic net (X1,Φ1) as in §6. We define the energy loss ratio of X0 with respect to A or At as

R(X0,Φ0,A(t))=E(X0,Φ0)E(A11(X1,Φ1))E(X0,Φ0).

If no local moves occur, then R=0. We may regard the ratio R as a degree of destruction. Note that R may be negative by some occurrence of contractions.

Consider the uniaxial extension by

A(λ)=diag(λ,λ1/(N1),,λ1/(N1))SL(N,R)

and At=A(λt). The permanent strain ϵ0 was defined in §4, by setting A11(X1,Φ1) as the reference position.

Proposition 8.2. —

Suppose that Φ0 is standard, no contractions occur in the deformation, and R=R(X0,Φ0,At)<1/N. Then, the permanent strain ϵ0 satisfies that

(1NN1R)(N1)/2N1ϵ0(1NR)(N1)/2N1.

Approximations

(1NN1R)(N1)/2N112R,(1NR)(N1)/2N1N12R,

hold when R0. One might expect that ϵ00 for λ>1, but it does not generally hold, because the directions of deformation and splittings do not necessarily coincide.

Proof. —

Write E0=E(X0,Φ0), E1=E(A11(X1,Φ1)) and (τij)=T(A11(X1,Φ1)). Then, T(X0,Φ0)=(E0/N)I, E1=τ11++τNN, and R=1E1/E0. Since only splittings occur, we have τiiE0/N for each 1iN by theorem 7.2. Hence,

τ11=E1(τ22++τNN)E1N1NE0.

Proposition 4.2 implies that

(1+ϵ0)2N/(N1)=τ22++τNN(N1)τ11=E1τ11(N1)τ11=1N1(E1τ111).

Since τ11E0/N, we have E1/τ11NE1/E0=N(1R). Since R<1/N and τ11E1((N1)/N)E0, we have

E1τ11(1N1N(1R))1=N(1R)1NR.

Therefore,

1NN1R(1+ϵ0)2N/(N1)11NR.

In the remainder of this section, we consider the simplest case of splitting; let X0 be a periodic graph such that only a single vertex exists in each period. We identify each i=(i1,,iN)ZN with a vertex of X0. Let ei denote the edge of X0 joining 0 and iZN. We write wi for the weight of ei. It is necessary that wi=wi. Let ρ:ZNRN be a period homomorphism, and let (u1,,uN) be a basis of ρ(ZN). The period homomorphism ρ induces a periodic realization Φ0 of X0 such that the image of vertices is ρ(ZN). For iZN, the edge ei corresponds to vi=i1u1++iNuNRN. Then, Φ0 is a harmonic realization. Let IZN such that ZN=II{0}. Let X1 be a periodic graph obtained from X0 by splitting the vertex 0 into v0 and v1 so that v0 and v1 are endpoints of ei for iI and iI, respectively, where ei is an edge of X1 obtained from ei. We write w0 for the weight of the new non-loop edge. Let Φ1 be a harmonic realization of X1 with the period homomorphism ρ. We may assume that Φ1(v0)=0. Let x=Φ1(v1).

For u0RN, let RuN={xRNux>0} and Iu=RuNZN. If IuI, the vertices split in the direction of u. Subsequently, it is possible to regard the net (X1,Φ1) as a result of slow deformation. However, we do not require this assumption unless otherwise stated.

After we show general behaviour, we give some examples of the energy loss ratio R=(E(X0,Φ0)E(X1,Φ1))/E(X0,Φ0) depending on the weights {wi}iZN.

Proposition 8.3. —

It holds that

x=iIwiviw0+iIwi

and

E(X0,Φ0)E(X1,Φ1)=||iIwivi||2w0+iIwi=(w0+iIwi)||x||2.

Proof. —

The condition of harmonic realization w0x+iIwi(vix)=0 implies the presentation of x. We have w010=w0, z=iIwivi and W=iIwi in the notation of theorem 7.2. Thus, we obtain the presentation of E(X0,Φ0)E(X1,Φ1).

We remark that Φ1 coincides with Φ(a) in theorem 7.6. Moreover, W attains the maximum in lemma 7.3.

By way of example, we suppose that the weights are given by a function of the lengths of edges.

Theorem 8.4. —

Let F(x) be a non-negative function on RN such that F(x)=F(x). Put wi=F(svi) for s>0. Suppose that the sum iIF(svi)||vi||2 is convergent. Let p=w0/w0, which is regarded as the probability that a loop changes into a non-loop edge. Suppose that there is u0RN such that IuI. Then, the energy loss ratio R(s,p) satisfies

lims0R(s,p)=||RuNF(x)xdx||2RuNF(x)dxRuNF(x)||x||2dx,

where we suppose that the three integrals are finite. Moreover, if F(0)>0 and limsiIF(svi)=0, then limsR(s,p)=0 for any fixed p>0.

By the second assertion, we may understand that the material has lower plasticity if the proportion of loops is large.

Proof. —

Since ||vi||1 for all but finitely many i, the sums iIF(svi) and ||iIF(svi)vi|| are absolutely convergent. The volume per period is given by V=det(u1uN). By proposition 8.3,

R(s,p)=||iIF(svi)vi||2(pF(0)+iIF(svi))iIF(svi)||vi||2.

Hence,

lims0R(s,p)=lims0||sNViIF(svi)svi||2(sNVpF(0)+sNViIF(svi))sNViIF(svi)||svi||2=lims0||sNViIF(svi)svi||2(sNViIF(svi))sNViIF(svi)||svi||2=||RuNF(x)xdx||2RuNF(x)dxRuNF(x)||x||2dx,

where the Riemann sums converge to the Riemann integrals.

Suppose that p>0, F(0)>0 and limsiIF(svi)=0. Since R(s,0)1, we have

limsR(s,p)=limsiIF(svi)pF(0)+iIF(svi)R(s,0)=0.

For example, we use the normal distribution, given by the function F(x)=(2πσ2)N/2exp(||x||2/2σ2) for xRN and σ>0. Then,

limσR(σ,p)=||Ru1Ne||x||2xdx||2Ru1Ne||x||2dxRu1Ne||x||2||x||2dx=(0ex2xdx(Rex2dx)N1)212(Rex2dx)NN2Rex2x2dx(Rex2dx)N1=4(0ex2xdx)2NRex2dxRex2x2dx=2Nπ,

and limσ0R(σ,p)=0 for p>0. Note that constant multiples of the weights do not change the energy loss ratio R. Generally, if the function F(sx) for each fixed x0RN monotonically decreases for s0 and iIF(svi)<, then limsiIF(svi)=0.

Next, we consider the weight functions given by the linear sums ws,i=(1s)w0,i+sw1,i for 0s1. The weight function ws,i for each s induces the energy

Es=E((X0,ws,i),Φ0)=iIws,i||vi||2=(1s)E0+sE1,

and the energy loss ratio Rs=||zs||2/W¯sEs by proposition 8.3, where

zs=iIws,ivi=(1s)z0+sz1

and

W¯s=pws,0+iIws,i=(1s)W¯0+sW¯1.

We consider the way in which Rs depends on the compounding ratio s. Since ||z0||2=W¯0E0R0 and ||z1||2=W¯1E1R1, we have

Rs=||(1s)z0+sz1||2W¯sEs=(1s)2W¯0E0R0+s2W¯1E1R1+2(1s)sz0z1W¯sEs.

Proposition 8.5. —

Suppose that R0=R1. Then

Rs=R0(1s)sW¯sEs((W¯0E1+W¯1E0)R02z0z1)R0.

The equality holds if and only if E0/W¯0=E1/W¯1 and the vectors z0 and z1 are parallel. For fixed w0,i and w1,i, the minimum of Rs is attained when

s=s^=W¯0E0W¯0E0+W¯1E1.

We may understand that a material with lower plasticity is obtained by blending two materials. Note that if I=Iu for u0RN, the vectors z0 and z1 are likely to be nearly parallel to u.

Proof. —

It is easy to check the presentation of Rs. The inequality RsR1 follows from

2z0z12||z0||||z1||=2W¯0E0W¯1E1R0(W¯0E1+W¯1E0)R0,

where the first inequality is the Cauchy–Schwarz inequality, and the second follows from (W¯0E1W¯1E0)20. An easy calculation shows that

(1s)sW¯sEs=(1s)s((1s)W¯0+sW¯1)((1s)E0+sE1)=s(1s)1(W¯0+s(1s)1W¯1)(E0+s(1s)1E1)

increases for 0<s<s^ and decreases for s^<s<1.

We remark that the linear sums of weights with different energy loss ratios do not yield smaller ratios in general. For instance, the linear sums of w0,i and ws^,i as in proposition 8.5 give the energy loss ratios between R0 and Rs^.

Under the assumption that R0=R1, we have

Rs^=R0+z0z1/W¯0E0W¯1E11+(W¯0E1+W¯1E0)/2W¯0E0W¯1E12R01+(W¯0E1+W¯1E0)/2W¯0E0W¯1E1

and

W¯0E1+W¯1E02W¯0E0W¯1E1=x2+12x,

where x=W¯0E1/W¯1E0. If x or x1 is large, then Rs^ is nearly equal to zero.

For example, we consider the cube lattice. Suppose that (u1,,uN) is the standard basis of RN. For mZ>0, let w0,0=w1,0=a, w0,±uk=w1,±muk=1 for any k=1,,N, and w0,i=w1,i=0 for the other i. Let u=(1,,1) and I=Iu. Then, W¯0=W¯1=pa+N, E0=N, E1=m2N, z0=(1,,1), z1=(m,,m) and R0=R1=1/(pa+N). Since s^=1/(1+m), we have Rs^=4m(1+m)2R0. Consequently, Es^=mN and limmRs^=0.

We give another example by using the linear sums of normal distributions. Suppose that wk,i=(2πσk2)N/2exp(||vi||2/2σk2) for k=0,1 and σk>0. Fix μ=σ1/σ0. Theorem 8.4 implies that

R~s=limσ0Rs=2(1s+sμ)2Nπ(1s+sμ2).

It attains the minimum when s=s^=1/(1+μ). Hence, R~0=R~1=2/Nπ, R~s^=4μ(1+μ)2R~0, and limμR~s^=0. We remark that limσ0W¯1/W¯0=1 and limσ0E1/E0=μ2.

9. Examples of deformation

In this section, we give examples of deformation for two-dimensional nets. Let

A(λ)=(cosθsinθsinθcosθ)(λ00λ1)(cosθsinθsinθcosθ).

We consider a uniaxial extension with strain ϵ=λ1 in the direction of an angle θ from the horizontal axis, namely, the slow deformation by At=A(λt).

For the splitting of a vertex v into v0 and v1, let e, e0,e1 and e01, respectively, denote the loops on v,v0,v1 and the non-loop edge between v0 and v1. Suppose that the weight function w satisfies w(e0)=w(e1)=w(e)/4 and w(e01)=w(e)/2.

Let X(m) denote the periodic graph after the mth local move, and let Φ(m) be a harmonic realization of X(m). Suppose that Φ(0) is standard, and all the period homomorphisms for Φ(m) are common. We write E(m)(λ)=E(A(λ)(X(m),Φ(m))). The engineering stress is σeng(m)(λ)=(1/V)(dE(m)/dλ). The permanent strain ϵ0(m) is the number satisfying σeng(m)(1+ϵ0(m))=0. The energy loss ratio is R(m)=1E(m)(1)/E(0)(1).

(a) . Hexagonal lattice

Let X(0) be a periodic graph of which the edges consist of loops and the 1-skeleton of the hexagonal tiling. We assume that the period is minimal. Let w0 and w1 be the weights of each loop and non-loop edge, respectively. A standard realization Φ(0) of X(0) is given as follows (figure 7):

  • (i)

    for l>0, the vectors u1=(3l,0) and u2=((3/2)l,(3/2)l) form a basis of the lattice;

  • (ii)

    representatives v0 and v1 of the vertices are mapped to Φ(0)(v0)=(0,0) and Φ(0)(v1)=((3/2)l,(1/2)l); and

  • (iii)

    The non-loop edges e1,e2 and e3 originating from v0 are mapped to Φ(0)(e1)=((3/2)l,(1/2)l),Φ(0)(e2)=((3/2)l,(1/2)l) and Φ(0)(e3)=(0,l).

Figure 7.

Figure 7.

The hexagonal lattice.

Then, the volume per period is V=(33/2)l2. The energy per period is E(0)=E(X(0),Φ(0))=3w1l2. Young’s modulus is E=4E0/V=(83/3)w1. The tension tensors around the vertex v0 and v1 are T(v0)=T(v1)=((3/2)w1l200(3/2)w1l2) .

(i) . The case θ=0

Consider the slow deformation by At=A(λt) for θ=0. Then,

E(0)(λ)=32w1l2(λ2+λ2),σeng(0)(λ)=233w1(λλ3).

Suppose that 1<δ1l<2K2w0+3w1/3w1l2. When λt=δ1l, the edge e3 contracts to a loop, where no splittings occurred earlier. We obtain a periodic graph X(1) as shown in figure 8. The edges of X(1) consist of loops with weight 2w0+w1 and the 1-skeleton of the square tiling. Then,

E(1)(λ)=32w1l2(λ2+3λ2),σeng(1)(λ)=233w1(λ3λ3)

and

ϵ0(1)=3410.316,R(1)=1.

Furthermore, when λt=K4w0+6w1/3w1l2, the vertices of X(1) split. We obtain a periodic graph X(2) as shown in figure 9. The weights of each loop and new non-loop edge are, respectively, (1/2)w0+(1/4)w1 and w0+(1/2)w1. Then,

E(2)(λ)=E(1)(λ)3w12l2w0+(5/2)w1λ2=32w1l2(2w0+w12w0+5w1λ2+3λ2),σeng(2)(λ)=233w1(2w0+w12w0+5w1λ3λ3)andϵ0(2)=3(2w0+5w1)2w0+w141,R(2)=2w0+3w12w0+5w1.

(The first equality is also obtained by theorem 7.2.) It holds that 341<ϵ0(2)15410.967. If w0=w1, then ϵ0(2)=7410.626. We remark that ϵ0(2) decreases as w0/w1 increases.

Figure 8.

Figure 8.

The graph X(1).

Figure 9.

Figure 9.

The graph X(2).

Furthermore, when λt=((2w0+5w1)/(2w0+w1))((2K2w0+3w1)/(3w1l2)), vertices of X(2) split. For genericity, we suppose that a single vertex per period splits. (We may assume that a representative thereof is at the origin.) We obtain a periodic graph X(3) as shown in figure 10. Then,

E(3)(λ)=E(2)(λ)2w0+w12w0+5w124w12l22w0+21w1λ2=32w1l2(2w0+w12w0+21w1λ2+3λ2),σeng(3)(λ)=233w1(2w0+w12w0+21w1λ3λ3)andϵ0(3)=3(2w0+21w1)2w0+w141,R(3)=2w0+11w12w0+21w1.

It holds that 341ϵ0(3)63411.817. If w0=w1, then ϵ0(3)=23411.189.

Figure 10.

Figure 10.

The graph X(3).

Similarly, more splittings of vertices may occur. Subsequently, the length of a path of edges increases. For w0=w1=1, the energies are shown in figure 11, and the stress–strain curve is drawn as the thick discontinuous curve in figure 12. Note that if we take a larger period, the occurrences of local moves change by genericity. We must also consider a contraction when λt=3δ1l. With this consideration, the stress–strain curve also changes. It would be desirable for these stress–strain curves to converge to a continuous curve as the periods expand.

Figure 11.

Figure 11.

Energies for deformation in the case θ=0.

Figure 12.

Figure 12.

Stress–strain curve in the case θ=0.

(ii) . The case θ=π/6

Consider the slow deformation by At=A(λt) for θ=π/6. Then, only splittings occur. We state only a result:

E(0)(λ)=32w1l2(λ2+λ2),σeng(0)(λ)=233w1(λλ3)( same as above) ,E(1)(λ)=32w1l2(3w03w0+4w1λ2+λ2),σeng(1)(λ)=233w1(3w03w0+4w1λλ3),ϵ0(1)=3w0+4w13w041,R(1)=2w13w0+4w1,E(2)(λ)=32w1l2(3w03w0+8w1λ2+λ2),σeng(2)(λ)=233w1(3w03w0+8w1λλ3),ϵ0(2)=3w0+8w13w041,R(2)=4w13w0+8w1,E(3)(λ)=32w1l2(w0w0+8w1λ2+λ2),σeng(3)(λ)=233w1(w0w0+8w1λλ3)andϵ0(3)=w0+8w1w041,R(3)=4w1w0+8w1,

and so on. For w0=w1=1, the energies are shown in figure 13, and the stress–strain curve is shown in figure 14.

Figure 13.

Figure 13.

Energies for deformation in the case θ=π/6.

Figure 14.

Figure 14.

Stress–strain curve in the case θ=π/6.

10. Discussion and perspectives

We discuss some open issues and propose questions for further research.

  • (1)

    The definition of the tension tensor fits with our purely mathematical interest, and we can apply it to topological and discrete geometric properties of graphs and nets. To define an energy, we have used E(l)=l2 as an energy of an edge of length l. What happens when we take another energy function E(l)? A linear transformation of a harmonic realization is no longer harmonic, but a harmonic realization is still unique if E(l)>0 and E(l)>0.

  • (2)

    In §5, we have arbitrarily given the threshold values δ and Kd in the conditions of a contraction and a splitting. What are the values δ and Kd for an actual TPE? How does the value Kd depend on the degree d?

  • (3)

    The validity of our model for deformation requires the assumption that infinite repetition of local moves does not occur. However, mutually inverse splittings and contractions may continue alternatingly in some cases. Theorem 6.1 gives a sufficient condition that such alternating repetition does not occur. Is another kind of infinite repetition possible? If it is possible, what condition is sufficient to avoid such repetition?

  • (4)

    In §9, we have given simple examples for two-dimensional nets. What about three-dimensional nets?

  • (5)

    Do local moves occur in unloading? If not, then it reproduces the Mullins effect [14]: the stress–strain curve in reloading coincides with that in the unloading until the maximal strain of the prior loading. Figures 12 and 14 illustrate this behaviour. Even if local moves occur in unloading, the inverse local moves in reloading may induce the Mullins effect.

  • (6)

    We have worked in continuous weights of edges, which is useful for mathematical arguments. Of course, the weights corresponding to actual polymers are integers. However, it would not be meaningful only to restrict ourselves to integer weights. Our presented settings preserve periodicity. As a result, stress–strain curves are not continuous. To obtain a continuum limit of such discrete matters, stochastic formulation of local moves will be effective. The stochastic formulation in integer weights may induce continuous stress–strain curves and a nonlinear constitutive equation that reflects elastoplasticity. We expect that this is also appropriate to describe fracture of materials.

Acknowledgements

The authors are grateful to Yoshifumi Amamoto, Tetsuo Deguchi, Ken Kojio, Motoko Kotani, Hiroshi Morita, Ken Nakajima, Genki Omori and Koya Shimokawa for their helpful discussions. The authors would like to thank Editage (www.editage.com) for English language editing, and the anonymous reviewers for their suggestions and comments.

Data accessibility

This article has no additional data.

Authors' contributions

K.Y. and H.K.: conceptualization, formal analysis, methodology, visualization, writing—original draft, writing—review and editing.

Both authors gave final approval for publication and agreed to be held accountable for the work performed therein.

Conflict of interest declaration

We declare we have no competing interests.

Funding

This study is supported by JST CREST grant no. JPMJCR17J4. The first author is also supported by the World Premier International Research Center Initiative (WPI), MEXT, Japan. The second author is also supported by JSPS KAKENHI grant no. 19K14530.

References

  • 1.Kotani M, Sunada T. 2000. Albanese maps and off diagonal long time asymptotics for the heat kernel. Commun. Math. Phys. 209, 633-670. ( 10.1007/s002200050033) [DOI] [Google Scholar]
  • 2.Kotani M, Sunada T. 2000. Jacobian tori associated with a finite graph and its Abelian covering graphs. Adv. Appl. Math. 24, 89-110. ( 10.1006/aama.1999.0672) [DOI] [Google Scholar]
  • 3.Kotani M, Sunada T. 2001. Standard realizations of crystal lattices via harmonic maps. Trans. Am. Math. Soc. 353, 1-20. ( 10.1090/S0002-9947-00-02632-5) [DOI] [Google Scholar]
  • 4.Kotani M, Sunada T. 2003. Spectral geometry of crystal lattices. Contemp. Math. 338, 271-306. [Google Scholar]
  • 5.Sunada T. 2012. Topological crystallography: with a view towards discrete geometric analysis, vol. 6. Surveys and Tutorials in the Applied Mathematical Sciences. Berlin, Germany: Springer. [Google Scholar]
  • 6.Fredrickson GH. 2006. The equilibrium theory of inhomogeneous polymers, vol. 134. International Series of Monographs on Physics. Oxford, UK: Oxford University Press. [Google Scholar]
  • 7.Treloar LRG. 1975. The physics of rubber elasticity. Oxford, UK: Oxford University Press. [Google Scholar]
  • 8.Aoyagi T, Honda T, Doi M. 2002. Microstructural study of mechanical properties of the ABA triblock copolymer using self-consistent field and molecular dynamics. J. Chem. Phys. 117, 8153-8161. ( 10.1063/1.1510728) [DOI] [Google Scholar]
  • 9.Morita H, Miyamoto A, Kotani M. 2020. Recoverably and destructively deformed domain structures in elongation process of thermoplastic elastomer analyzed by graph theory. Polymer 188, 122098. ( 10.1016/j.polymer.2019.122098) [DOI] [Google Scholar]
  • 10.Liu H, Liang X, Nakajima K. 2020. Direct visualization of a strain-induced dynamic stress network in a SEBS thermoplastic elastomer with in situ AFM nanomechanics. Jpn J. Appl. Phys. 59, SN1013. ( 10.35848/1347-4065/ab948a) [DOI] [Google Scholar]
  • 11.Wells AF. 1977. Three-dimensional Nets and Polyhedra. Wiley monographs in crystallography. New York, NY: Wiley-Interscience. [Google Scholar]
  • 12.Gurtin ME. 1981. An introduction to continuum mechanics. New York, NY: Academic Press. [Google Scholar]
  • 13.Ericksen JL. 2008. On the Cauchy–Born rule. Math. Mech. Solids 13, 199-220. ( 10.1177/1081286507086898) [DOI] [Google Scholar]
  • 14.Diani J, Fayolle B, Gilormini P. 2009. A review on the Mullins effect. Eur. Polym. J. 45, 601-612. ( 10.1016/j.eurpolymj.2008.11.017) [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

This article has no additional data.


Articles from Proceedings. Mathematical, Physical, and Engineering Sciences are provided here courtesy of The Royal Society

RESOURCES