You are on page 1of 23

Acta Mech 234, 2963–2984 (2023)

https://doi.org/10.1007/s00707-023-03545-y

O R I G I NA L PA P E R

Yucheng Zhou · Kefu Huang

A simplified deformation gradient theory and its


experimental verification

Received: 1 December 2022 / Revised: 1 February 2023 / Accepted: 6 March 2023 / Published online: 28 March 2023
© The Author(s), under exclusive licence to Springer-Verlag GmbH Austria, part of Springer Nature 2023

Abstract In this paper, materials with nonlocal properties are considered as a continuum model composed of
micro-elements with certain volumes. Based on this hypothesis, the deformation and corresponding energy of
micro-structure system are studied in detail, and the equivalent governing equations in simplified form are given.
In the framework of micro-structure system, a simplified deformation gradient theory (SDG) with two length-
scale parameters is obtained by defining the micro-strain and micro-rotation of elements specifically from the
perspective of deformation, which has definite physical significance. The generalized strain energy is introduced
in the SDG, which gives a new explanation of elastic moduli, and the nonlocal effect parameter is defined
to capture nonlocal properties of materials quantitatively. Under certain micro-deformation assumptions, the
SDG can be degenerated into couple stress theory, strain gradient theory and classical continuum theory. The
nonlocal deformation consists of two branches: one is the macro-deformation of non-uniform materials and the
other is the micro-deformation of materials with micro-structures. For the macro-tension of particle reinforced
composites, the SDG successfully verifies and predicts the approximatively linear relationship between elastic
moduli and particle sizes on the micron scale. Moreover, the theoretical solution to nonlocal micro-torsion
based on the SDG agrees well with the experiment results and also predicts the torsion stiffness of cylinder at
smaller diameters.

1 Introduction

There are many size-dependent (or nonlocal) experiment phenomena under different deformation models,
including vibration and bending of micro-beams [1–4], bending of micro-plates [5], torsion and bending of
foam materials [6,7], plastic deformation of metal materials or epoxy resins [8,9] and macro-deformation
of particle reinforced composites [10–13]. The classical local continuum theory assumes that a particle only
interacts and exchanges energy with neighboring particles, which causes the lack of length-scale parameters
that are applicable to different scale models, and only valid for materials with long wavelengths. Therefore,
Eringen et al. [14], Krner [15], Kunin [16] and Silling [17] introduced different types of nonlocal continuum
theories to deal with the nonlocal effect problems.
Nonlocal theories can be classified into two major types, one is the integral theories and the other is
differential theories. The governing equations of integral theories are expressed by displacement fields rather
than their derivatives [18,19], which can be applied to deal with singular problems. Kunin [16] derived a three-
dimensional nonlocal model by approximating the continuum to a discrete lattice structure. Peridynamics
theory proposed by Silling [17] is more general since it considers two- and three-dimensional medium in
addition to one-dimensional medium. However, integral nonlocal theories weaken basic physical quantities

Y. Zhou · K. Huang (B)


Department of Mechanics and Aerospace Engineering, Southern University of Science and Technology, 1088 Xueyuan Boulevard,
Shenzhen, China
E-mail: huangkf@sustech.edu.cn

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


2964 Y. Zhou, K. Huang

such as stress and strain, which does not conform to the theoretical form of traditional mechanics. Thus,
various forms of differential nonlocal theory are also proposed [20–23]. The differential nonlocal theories apply
displacement gradient, displacement curl and their higher-order derivatives as well as introducing length-scale
parameters to construct constitutive equations. Scholars believe that, length-scale parameters in differential
theories are related to the micro-structure of materials [22,24]. Therefore, in some mechanical practices,
differential nonlocal theories can also explain size effect phenomena of materials.
In couple stress theories, particles in a continuum not only translate but also rotate under external loadings,
leading to the work done by both stress to strain and couple stress to rotation appears in the system deformation
energy. The prototype of couple stress theory was first proposed by Cosserat [25], and the Cosserat theory
containing couple stress was established. Prior to this, Voigt [26] proposed the concepts of body couple and
surface couple, and constructed a continuous model considering the couple stress acting on the surface or
boundary of a body. It was not until the 1960 s that some scholars began to modify and expand the Cosserat
theory. By imposing certain constraints on the rotation of material particles in the Cosserat continuum theory,
they gradually developed a more general theory, i.e., the couple stress theory [27–29].
For isotropic linear elastic materials, there are three additional parameters related to the micro-structure
of material besides two classical Lamé coefficients in the constitutive relations of general couple theory. By
introducing higher order equilibrium equations, a modified couple stress theory is proposed by Yang et al [30].
The constitutive equations of Yang’s theory contain only one additional parameter, which greatly reduces the
difficulty of determining the additional micro length-parameter. Park et al. [31] applied this modified couple
stress theory to calculate the bending of Euler-Bemoulli micro-beams, and found that micro-beams show
obvious size effect when the thickness is equal to the characteristic scale of materials. Based on couple stress
theories, Shaat [32] made a physical discussion and mathematical representation of the couple stress effect of
micro/nano solids in linear elastic materials, and evaluated the effectiveness and applicability. Neff et al. [33]
reviewed some fundamental misunderstandings in the indeterminate couple stress model. In recent years, the
couple stress theories have been applied to many researches, including elastic plates [34,35] and other different
fields [36–39].
Strain gradient theories construct new governing equations by introducing additional terms related to strain
gradient to the constitutive equation and strain energy of system. Mindlin [20] proposed a second order strain
gradient theory by considering the strain energy as a function of the strain and its first and second order
gradients. By degrading the second order strain gradient theory, Mindlin [21] developed a more widely used
first order strain gradient theory, which contains 18 material parameters and is a complete strain gradient theory.
Subsequently, various strain gradient plasticity theories were established since 1980 s. Aifantis [40] proposed
a strain-gradient plasticity theory by using the first and second Laplacian operators of equivalent strains to
represent additional strain gradients. Fleck et al. [41] and Han et al. [42] developed another strain-gradient
plasticity theory based on geometric essential dislocations. Aifantis [43] established and developed various
strain gradient theories to simulate the elastic, plastic and dislocation dynamic behaviors of nonlocal materials.
Fleck et al. [41,44] re-expressed Mindlin’s strain gradient theory by decomposing the second order defor-
mation gradient tensor into two independent parts, namely the tension gradient tensor and rotation gradient
tensor, which makes the deformation energy has a clear physical meaning. Inspired by the work of Yang
[30], Lam et al. [45] redefined the high-order strain tensor and its conjugate high-order stress tensor after
applying the additional coupling moment equilibrium relations. Lam et al. [45] further studied the bending
problem of micro-cantilever beams and found that the dimensionless stiffness of micro-beams is quadratic
inversely proportional to thickness. Zhou et al. [46] redefined the elastic theory of isotropic strain gradient
with five high-order elastic constants and proved that the number of independent high-order elastic constants of
isotropic materials was three. Fu et al. [47] compared different forms of nonlocal theories and evaluated them
with specific examples. Rahimi et al. [48] proposed a generalized nonlocal stress–strain gradient theory using
fractional calculus, which can be reduced to a variety of typical nonlocal theories. Fu et al. [49] established a
size-dependent model of Euler–Bernoulli beams based on the strain gradient theory and surface elastic theory,
in which the strain gradient parameter was used to capture strain gradient effect. There are many further studies
and practical applications based on strain gradient theories, one may refer to related papers [50–54].
Nonlocal theories mostly introduce length-scale parameters by modifying constitutive equation of clas-
sical continuum theory and adding higher-order displacement gradient terms to meet the nonlocal properties
of materials. The couple stress theories only add rotation gradient that is conjugated with couple stress in
constitutive equation, while the strain gradient theories consider the effects of tensile gradient and expansion
gradient, or maybe also rotation gradient in some cases. Although these two theories can explain part of size
effect and singularity problems, their constitutive relations have no clear deformation significance. In order to

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


A simplified deformation gradient theory 2965

explain deformation characteristics of differential nonlocal theories including couple stress theory and strain
gradient theory, Mindlin [55] proposed a complete framework of micro-structure theory.
In this paper, the material with nonlocal properties is regarded as a continuum composed of microscopic
elements and a simplified micro-structural deformation gradient theory is constructed based on the deforma-
tion of micro-elements. Under external loadings, micro-strain and micro-rotation will be generated in addition
to macro-deformation in non-uniform materials or micro-materials. The micro-structure system is then con-
structed and the deformations and corresponding energy expressions are studied in detail. Defining the strain
and rotation of micro-elements separately and associating these two parts of micro-deformation with macro-
strain and macro-rotation, we obtain a concise differential nonlocal theory with two length-scale parameters,
namely the SDG. Without losing scientificity, the SDG greatly reduces the number of length-scale parame-
ters through reasonable micro-deformation assumptions, and can be reduced to a series of nonlocal theories,
including classical couple stress theory and strain gradient theory, which has definite physical meanings and
significant application values. Based on the conservation of deformation energy and the definition of gener-
alized strain energy, we reinterpret the physical meaning of elastic moduli. The introduction of length-scale
parameters related to micro-structure of material can be used to quantitatively capture various size effects
and solve singularity problems. We apply the SDG to nonlocal deformations, and theoretical solutions to the
macroscopic uniaxial tensile size effect of particle reinforced composites and the micro-torsional size effect
of cylinders are obtained, which are in good agreement with experiment results.

2 Micro-structure system and its nonlocality

2.1 Basic concepts in the micro-structure system

In classical continuum theory, locality is one of the most fundamental assumptions, which plays an important
role in simplifying constitutive equations. The classical continuum theory assumes that material is composed
of infinitesimal particles, and the stress of any particle is only related to the particles in neighborhood, i.e., the
local action principle. The particles in material cannot be subjected to moments, but only stresses. Therefore,
there is equivalent law of shearing stresses exists in classical continuum theory, and the definition of “simple
material” is given based on such properties.
However, the micro-structure of material is often not negligible in many specific problems. This is especially
obvious when there is little difference between the macroscopic and microscopic structural dimensions of
materials. Thus, we cannot simply consider the material to be composed of particles, but must examine the
influence of micro-structure on the mechanical properties. On the other hand, when the object is composed
of non-uniform material, such as ferroelectric materials with grains, particle reinforced composite materials
with inclusions, etc., the nonlocal effect brought by non-uniform medium on the whole material should also
be studied. Therefore, we may define the material with micro-elements that can be subjected to both forces
and moments as the "actual material". The materials in the above two cases are two representative types of
actual material, and the equivalent law of shearing stresses is not necessarily applicative here.
(i) Double force
Double force is a force system of two opposite forces with the same magnitude and can be represented by
a second order asymmetric tensor  = Φi j . The first subscript of double force tensor gives the orientation
of lever arm ri between two forces and the second subscript denotes the orientation of forces F j . There are
no further cross product operations between forces and lever arms in the definition of double forces. The
simplified principal vector of double force must be zero based on the simplification rules for force systems,
thus the components of double force tensor essentially denote a couple vector or a self-equilibrating pair of
forces, as shown in Fig. 1.
The couple vector corresponding to double force tensor is denoted by symbol [·], to be specific,

[] = [Φi j ] ≡ ri ei × F j e j = − ×
• I, (1)

where × and • are symbols for the cross and dot products respectively in tensor operations, and the combination
of them satisfies the following operation law,
× ×
 
Φ • I = Φi j ei e j • ek ek = Φi j (ei · ek ) e j × ek . (2)

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


2966 Y. Zhou, K. Huang

Fig. 1 Typical components of double force or double force per unit area (double stress)

The couple vectors of the symmetric and anti-symmetric parts of double force tensor are denoted by [S ]
and [A ] respectively, i.e., [] = [S ] + [A ]. By virtue of Eq. (1), we have

[S ] = −S ×
• I = 0, (3)
which indicates that the symmetric part of double force produces no moments and is a self-equilibrating pair
of forces. Similar conclusions can also be referred to the paper [24].
The couple vector of anti-symmetric part of double force satisfies
×
[A ] = −A • I  = 0. (4)
That is to say, the anti-symmetric part of double force is not self-equilibrating and has effect of a couple
moment, whence we may define the anti-symmetric part of double force tensor as couple force tensor. Since
any second order anti-symmetric tensor is dual to an axial vector, the dual vector of A is also a couple moment
and satisfies
1 A × 1
Φ=  • I = − [A ]. (5)
2 2
Equation (5) shows that the dual vector and couple vector of anti-symmetric A are collinear vectors in
opposite directions, and the length of couple vector is twice that of dual vector.
(ii) Double stress
The double stress is a third order tensor with 27 components, denoted as μ = μi jk ≡ ∂i Φ jk . The first
subscript of the double stress tensor indicates the normal direction of surface on which double force applies.
The second and third subscripts have the same significance as two subscripts of double force tensor. Similar
to the way double force tensor is decomposed, we can also resolve double stress tensor into symmetric and
anti-symmetric parts, μ = μS + μA , satisfying
1  1 
μS = μi( jk) ≡ μi jk + μik j , μA = μi[ jk] ≡ μi jk − μik j . (6)
2 2
To be specific, the symmetric part of double stress arises from double force without moment, i.e., the symmetric
part of double force tensor. The symmetric part of the double stress tensor is also self-equilibrating as the
symmetric part of double force tensor [55]. The anti-symmetric part of double stress arises from double force
with moment, i.e., the anti-symmetric part of double force tensor, which is defined as the couple stress tensor.
Since the couple force is essentially a couple vector, the couple stress is a second order tensor with nine
components.

2.2 Micro-structure system

2.2.1 Definition of generalized strain energy

In the micro-structure system, the material is composed of micro-elements with a certain volume, and the
elements can be subjected to both forces and moments. In the micro-structure system, the deformation of
macro-body is often inconsistent with that of micro-elements, thus two sets of macroscopic and microscopic
coordinate systems are adopted. The macroscopic coordinate system xi (i = 1, 2, 3) is established on the
macro-body, while the microscopic coordinate system xi (i = 1, 2, 3) is a fixed coordinate system on the
micro-elements. Denote the spatial position vectors of particles in the macroscopic and microscopic coordinate

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


A simplified deformation gradient theory 2967

Table 1 Definitions of basic physical quantities in the micro-structure system [55]

Basic physical quantities Definition based on displacement


 
Macro-strain εi j = ∂i u j + ∂ j u i /2
Macro-rotation ωi j = ∂i u j − ∂ j u i /2
Micro-strain ψ(i j) =  ψi j + ψ ji  /2
Micro-rotation ψ[i j] = ψi j − ψ ji /2
Relative deformation γi j = ∂i u j − ψi j
Micro-deformation gradient i jk = ∂i ψ jk

systems before and after deformation by X i , X i and xi , xi respectively, then the macroscopic and microscopic
displacements are defined as
u i ≡ xi − X i , u i ≡ xi − X i (7)
Then, the deformation of macro-body is called macro-deformation ∂i u j , and the deformation of micro-elements
is micro-deformation ψi j ≡ ∂i u j , where ∂i (·) and ∂i (·) are derivatives with respect to the macro coordinate
xi and micro coordinate xi respectively. The difference between micro-deformation and macro-deformation
is defined as the relative deformation γi j ≡ ∂i u j − ∂i u j . The definitions of basic mechanical quantities in
micro-structure system are given in Table 1.
Mindlin [55] derived the stress and double stress equilibrium equations of micro-structure systems applying
the energy method,
∇ · (τ + σ ) + f = 0, ∇ · μ + σ +  = 0, (8)
where τ , σ and f are Cauchy stress, relative stress and body force per unit volume, μ and  are double stress
tensor and double force tensor per unit volume. The traction and moment boundary conditions are
t = n · (τ + σ ) , T = n · μ, (9)
where t is the surface force per unit area and T is double force per unit area. The anti-symmetric part of double
force  is the body couple, and the anti-symmetric part of double traction T is the Cosserat couple stress
vector.
Based on Hamilton’s principle, we obtain the virtual work principle of micro-structure system,
δ = δ U − δ W = 0, (10)
where δ U is the variation of total strain energy, and δ W is the variation of work done by external loadings,
and δ is then defined as the variation of total potential energy of the deformable body.
Due to the existence of micro-elements, the work done by external loadings in micro-structure system
includes the work of surface couples and body couples besides the work done by body forces and surface
forces in classical continuum theory. Therefore, for an object consisting of micro-elements, the variation of
total work done by external loadings is
   
δW = f · δu dV +  : δψ dV + t · δu dS + T : δψ dS, (11)
∂ ∂

where δu and δψ are the variation of macro-displacement and micro-displacement gradient respectively.
.
Substituting Eqs. (8) and (9) into Eq. (11), and considering ∇ · (μ : δψ) = (∇ · μ) : δψ + μ .. (δψ) ∇,
we find
  
.
δW = τ : (δu) ∇ + σ : [(δu) ∇ − δψ] + μ .. (δψ) ∇ dV, (12)

and the total work done by external loadings can be further written as
  
.
δW = τ : δε + σ : δγ + μ .. δκ dV. (13)

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


2968 Y. Zhou, K. Huang

Fig. 2 Deformation and corresponding energy of micro-structure system

By virtue of the principle of virtual work Eq. (10) and from Eq. (13), we have
.
δU = δW = τ : δε + σ : δγ + μ .. δκ, (14)

where δU is the variation of generalized strain energy density, and δW denotes the variation of external works
per unit volume. The first term represents the variation of total strain energy density δUs , which is produced
by the classical symmetric strain. Due to the existence of relative deformation, the deformations of micro-
elements and macro-body are inconsistent, which leads to the generation of relative energy density Ur . The
last term is the variation of high-order gradient energy density δUg , including deformation energy generated
by strain gradient and rotation gradient. The detailed deformation and corresponding deformation energy of
the micro-structure system are given in Fig. 2.
(i) The strain energy density Us
Under the assumption of simple material, any deformation resulting from symmetric stresses can be described
by the superposition of simple tensile and simple shear. In elastomer , the element occupies abcd in unde-
formed state has pure shear deformation under shear stress, and deformed to a  b c d  . Under the action of
normal stress, the element expands into a  b c d  , as shown in Fig. 3a. To be specific, the strain energy density
in terms of classical strain components is
1 1
Us = τi j εi j = λεii ε j j + μεi j εi j , (15)
2 2
where λ and μ are elastic constants having the same properties as, but slightly different from, the classical
Lamé coefficients. According to the deformation and energy in micro-structure system mentioned above, we
know that the external work is considered to be also converted into other forms of deformation energy besides
classical strain energy in the micro-structure system, while the external work based on the classical theory is
all converted into classical strain energy, which leads to the difference between the elastic constants λ and μ
and classical Lamé coefficients.
(ii) The relative deformation energy density Ur
In the classical continuum theory, the rotation deformation of macro-body (i.e., the anti-symmetric part of
macro-displacement gradient) is often negligent in the study of strain energy, since macro-rotation does not
produce change of deformation energy. However, in micro-structure system, the rotation of micro-elements
will cause deformation of macro-body, resulting in the relative deformation and corresponding rotation related
energy, as shown in Fig. 3b.
The relative stress σ and relative strain γ defined in the micro-structure system are both asymmetric.
Decomposing the relative stress and relative strain into symmetric and anti-symmetric parts, we have
1 1 S    1 1
Ur = σ :γ = σ + σ A : γ S + γ A = σ S : γ S + σ A : γ A. (16)
2 2 2 2

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


A simplified deformation gradient theory 2969

Fig. 3 a Dilatation and shear deformation of symmetric stress. b Rotation deformation of anti-symmetric stress

Equation (16) shows that the relative deformation energy includes two parts, namely the energy produced by
symmetric and anti-symmetric parts of relative strain. Define
1 S 1
UrS ≡ σ : γ S , UrA ≡ σ A : γ A . (17)
2 2
The symmetric and anti-symmetric parts of the relative deformation correspond to the difference of strain and
rotation between the micro-element and macro-body respectively. Symmetric stress produces dilatation and
shear deformation, while anti-symmetric stress produces rotation deformation, as shown in Fig. 3.
(iii) The high-order gradient energy density Ug
Decompose the micro-deformation into symmetric part and anti-symmetric part, ψ = ψ S +ψ A , the symmetric
part of micro-deformation denotes micro-strain, and the anti-symmetric part denotes micro-rotation, one can
refer to Table 1.
The micro-deformation is only a function of macroscopic position coordinates in this paper. The micro-
deformation gradient can be also decomposed into two parts, i.e., the macroscopic gradient of micro-strain
and the macroscopic gradient of micro-rotation,
κ ≡ ∇ψ = ∇ψ S + ∇ψ A . (18)
The corresponding high-order gradient energy density Ug may also be decomposed into macroscopic gradient
energy density of micro-strain and micro-rotation,

Ug = Ug(S) + Ug(R) . (19)


The total elastic deformation energy in micro-structure system is defined as the generalized strain energy,
U ≡ Us + Ur + Ug , (20)
where Us is the strain energy density, generated by symmetric macro-strain and Cauchy stress. Ur and Ug are
the relative deformation energy density and high-order gradient energy density respectively. These two parts
of deformation energy are related to the deformations of micro-elements, which can also be called nonlocal
energy. The treatment of each part of generalized strain energy in the SDG will be studied in detail in Sect. 3.1.

2.2.2 Relative deformation and relative stress

The first of Eq. (8) gives the stress (force) equilibrium equation, and the second represents the double stress
(moment) equilibrium equation. The relative stress σ appears in both stress equilibrium equation and double
stress equilibrium equation. In order to examine the physical meaning of relative stress, we define asymmetric

total stress τ = τ + σ , then the stress equilibrium equation and traction boundary condition become
◦ ◦
∇ · τ + f = 0, t = n · τ . (21)

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


2970 Y. Zhou, K. Huang

The moment equilibrium equation of the micro-structure system is


  
  
 dV + r × f dV + T + r × t dS = 0, (22)

where  and T are body couple and surface couple vector. By virtue of the inverse commutativity of cross
product of vectors and commutativity of the mixed product of vectors, we have the moment equilibrium
equation,
◦ ×
 − τ • I + ∇ · μ = 0, (23)
μ
where is second order couple stress tensor and I is second order unit tensor.
In micro-structure system, the double force and couple force are both represented by a second order tensor,
and the specific operation is to separate the lever arm and forces without cross production. Along the same
lines, a second order tensor is applied to define the result of cross product of arbitrary vectors a and b without
doing the final operation using operator ⊗ in this part, i.e., a ⊗ b = ai b j , then we obtain
◦  A T
τ × ×
• I=σ • I≡σ • I= σ
A ⊗
, (24)

where ”T” denotes transpose of a tensor. Decompose the relative stress as σ = σ S +σ A , and τ × S×
• I = σ • I = 0.
 A T
Taking into account σ = −σ , the moment equilibrium equation becomes
A

A + σ A + ∇ · μA = 0. (25)
Equation (25) indicates that the symmetric part of relative stress is self-equilibrating and produces no moments.
Only the anti-symmetric part of the relative stress actually contributes to the moment equilibrium.

2.2.3 Equivalent form of basic equations

The general theories of micro-structure system are comprehensive nonlocal theories, which are of great impor-
tance due to their wide coverage and physical models that can better perform actual materials. However, too
many material parameters in these theories make it difficult to deal with practical problems. Therefore, it is
necessary to simplify physical quantities as much as possible, and give the simplified equivalent forms of
equilibrium equations and boundary conditions, which play an important role in solving practical problems.
We also decompose the total stress (or equilibrium stress) into symmetric and anti-symmetric parts,
◦ ◦ S ◦ A  
τ= τ + τ = τ + σ S + σ A, (26)

where tensor σ A is the anti-symmetric part of both tensors σ and τ . On account of the physical meaning
of symmetric part and anti-symmetric part of double stress, we may give the equivalent form of equilibrium
equations,
◦ ◦ A
∇ · τ + f = 0, ∇ · μA + τ + A = 0, (27)

and the equivalent form of boundary conditions,



t = n · τ , T = n · μA , (28)
where the symmetric part of σ , μ and  are self-equilibrating, thus they cannot appear in the moment equi-
librium equations, which plays an important role in simplification of the balance of micro-elements.
Remark (i) Mindlin obtained the equilibrium equations and traction boundary conditions from the perspective
of deformation energy, while we derive the equilibrium equations more directly from loadings on the body,
which have more intuitive physical significance.
(ii) The self-equilibrating force system does not contribute to moment equilibrium and double stress
boundary conditions, but it can generate the deformation and generalized strain energy of elements. Therefore,
the basic equations obtained by Mindlin based on energy contains the self-equilibrating terms, which is not
inconsistent with the simplified equilibrium equations.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


A simplified deformation gradient theory 2971

Table 2 The elastic variables based on displacement components

Elastic variables Definition based on displacement


 
Classical macro-strain εi j = ∂i u j + ∂ j u i /2 = ε ji
Classical macro-rotation ωi j = ∂i u j − ∂ j u i /2 = −ω ji
First gradient of classical strain κ̂i jk = ∂i ε jk = κ̂ik j
First gradient of classical rotation κ̄i jk = ∂i ω jk = −κ̄ik j
Second gradient of displacement κ̃i jk = ∂i ∂ j u k = κ̂i jk + κ̄i jk = κ̃ jik

3 Construction of the simplified deformation gradient theory

Compared with the general micro-structure theory, the SDG given in this section makes more reasonable and
specific assumptions on the micro-deformation, and has a wider application range without losing the premise
of scientific nature.

3.1 Construction of the simplified deformation gradient theory

3.1.1 Basic assumptions

In a micro-structure system, if the strain and rotation generated by the micro-elements and the macro-body
are synchronized, then the classical theory can be adopted to the mechanical analysis of these kind of simple
materials. In the case where the deformations of the micro-elements are different from that of the macro-body,
then we need to apply the modified continuum theory. Based on this fact, we introduce physical quantity α to
describe the proportional relationship between the strain of the micro-elements and the strain of the macro-
body. In this way, the mechanical properties of materials with micro-structures can be described relatively
accurately and simply.
As a modification of the classical continuum theory, micro-structure system is a phenomenological elastic
model describing mechanical behaviours of actual material. Relative rotation is defined as the difference
between rotation components of micro-elements and macro-body. In actual experiments, it is relatively difficult
to measure accurate micro-rotation components, and the experimental results show that the rotation of micro-
elements does exist under certain deformation conditions. Therefore, we associate the micro-rotation gradient
with the macro-rotation gradient by assuming that the relative rotation components are approximatively uniform
in the macroscopic coordinate system.
The micro-deformation is considered as combination of micro-strain and micro-rotation, corresponding
to symmetric part and anti-symmetric part of the micro-deformation respectively, which is similar to the
decomposition of macro-deformation. The basic mechanical quantities in the SDG defined by displacement
components and higher-order strain are given in Table 2.
In the SDG, the specific micro-deformation hypotheses are as follows:
(i) Micro-strain deformation hypothesis: the strain of micro-elements is proportional to the strain of macro-
body, i.e., ψ(i j) = αεi j , where α is the proportional coefficient of micro-strain, and 0  α  1 is satisfied.
Then, the relative deformation γ(i j) = (1 − α)εi j , and the macroscopic gradient of micro-strain is

κi( jk) = ∂i ψ( jk) = α∂i ε jk = α κ̂i jk . (29)

(ii) Micro-rotation deformation hypothesis: the macroscopic gradient of relative rotation is zero, i.e.,
∂i γ[ jk] = 0. Since the micro-deformation ψ jk is a function of only macroscopic position, the micro-rotation
gradient satisfies
 
κi[ jk] = ∂i ψ[ jk] = ∂i ω jk − γ[ jk] = ∂i ω jk = κ̄i jk . (30)

The preceding two independent deformations, the micro-strain and the micro-rotation, constitute a complete
micro-deformation hypothesis in the SDG.
The relative deformation energy density can be decomposed into relative strain energy density and relative
(S) (R)
rotation energy density based on the hypotheses of micro-strain and micro-rotation, i.e., Ur = Ur + Ur ,
then

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


2972 Y. Zhou, K. Huang

1
Ur(R) = (d2 − d3 ) γ[ jk] γ[ jk] + αd2 ε jk γ[ jk] ,
2
1 1
Ur(S) = (1 − α)2 d1 ε j j εkk + (1 − α)2 (d2 + d3 ) ε jk ε jk . (31)
2 2
where d1 − d3 are constant material parameters.
Accordingly, κi jk = ∂i ψ( jk) + ∂i ψ[ jk] reduces to
1 1 1 1
κi jk → α(κ̃i jk + κ̃ik j ) + (κ̃i jk − κ̃ik j ) = (1 + α)κ̃i jk − (1 − α)κ̃ik j . (32)
2 2 2 2
Thus, the high-order gradient energy density in terms of κ̃i jk is

Ug = ã1 κ̃iik κ̃k j j + ã2 κ̃i j j κ̃ikk + ã3 κ̃iik κ̃ j jk + ã4 κ̃i jk κ̃i jk + ã5 κ̃i jk κ̃k ji , (33)
where ã1 − ã5 are constant material parameters.
Following the definitions in Table 2 and eliminating displacement components, we obtain
1 
κ̂i jk = κ̃i jk + κ̃ik j , κ̄i jk = κ̂ jki − κ̂ki j . (34)
2
Applying κ̃i jk = κ̂i jk + κ̄i jk and Eq. (33), we split Ug into strain gradient energy expressed by macro-
strain gradient and rotation gradient energy expressed by macro-rotation gradient to describe the energy
corresponding to micro-strain and micro-rotation more directly and clearly,

Ug(S) = â1 κ̂iik κ̂k j j + â2 κ̂i j j κ̂ikk + â3 κ̂iik κ̂ j jk + â4 κ̂i jk κ̂i jk + â5 κ̂i jk κ̂k ji , (35)
and
Ug(R) = ā1 κ̄iik κ̄k j j + ā2 κ̄i j j κ̄ikk + ā3 κ̄iik κ̄ j jk + ā4 κ̄i jk κ̄i jk + ā5 κ̄i jk κ̄k ji , (36)

where â1 − â5 and ā1 − ā5 are constant material parameters. Of course, Eq. (36) can be expressed by Eq. (35).
We first simplify the relationship between strain gradient deformation energy and macro-strain gradient.
Considering the symmetry of strain in deformation [24], we select the coefficients â1 − â5 in Eq. (35) as
1 2  2
â2 = α λ  , â4 = μ α 2 2 , â1 = â3 = â5 = 0, (37)
2
where  is the strain gradient length-scale parameter, and there remain ε jk and κ̂i jk in Eq. (35), hence
1 2  2
Ug(S) = α λ  κ̂i j j κ̂ikk + α 2 μ 2 κ̂i jk κ̂i jk . (38)
2
It must be noted here that α and  can be multiplied as a material parameter, but since the definition of α in
deformation hypothesis has clear physical meaning, we still keep α and  separately, which is also convenient
to analyze the influence of parameters in subsequent studies.
Next we study the characteristics of rotation gradient deformation energy and its simplification. Since ωi j
is an anti-symmetric tensor, then ωii = 0, and the first two terms in Eq. (36) vanish automatically. Or we may
use the classical expression of couple stress theory to reexpress the third order rotation gradient tensor by the
gradient of rotation vector, which is a second order tensor, namely the curvature tensor χi j . The macro-rotation
vector wi and the second order curvature tensor χi j are defined as
1
wi = ei jk ω jk , χi j = ∂i w j , (39)
2
where ei jk is the alternating tensor, thus we have
x̄i jk x̄i jk = 8χi j χi j ,
 
x̄iik x̄ j jk = 4(1 − δi j ) χi j χi j − χi j χ ji ,
 
x̄i jk x̄k ji = 4(1 − δi j ) χi j χi j − χii χ j j . (40)

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


A simplified deformation gradient theory 2973

(R)
Therefore, the rotation gradient energy density Ug can be equivalently expressed by the curvature tensor,

Ug(R) = ā1 χi j χi j + ā2 χi j χ ji + ā3 χll2 . (41)


We have proved that two forms of rotation gradient energy expressed by Eqs. (36) and (41) are equivalent.
Yang et al. [30] conclude that only the symmetric part of displacement gradient (conventional strain tensor)
and the symmetric part of rotation gradient (symmetric curvature tensor) contribute to the deformation energy.
The rotation vector and the anti-symmetric curvature tensor do not contribute to deformation energy. Therefore,
rotation gradient energy density becomes

Ug(R) = μ L2 χi j χi j , (42)


where L is the rotation gradient length-scale parameter.
By virtue of ε jk γ[ jk] = 0 and Eq. (31), we obtain the relative deformation energy density in the SDG,

Ur = Ur(S) + Ur(R)
1 1 1
= (1 − α)2 d1 ε j j εkk + (1 − α)2 (d2 + d3 ) ε jk ε jk + (d2 − d3 ) γ[ jk] γ[ jk] . (43)
2 2 2
Substituting Eqs. (38) and (41) into Eq. (19), we obtain the high-order gradient energy density of the SDG,
1 2  2
Ug = Ug(S) + Ug(R) = α λ  κ̂i j j κ̂ikk + α 2 μ 2 κ̂i jk κ̂i jk + μ L2 χ jk χ jk . (44)
2

3.1.2 Governing equations

(i) The generalized strain energy density


Based on the deformation hypotheses in Sect. 3.1 and the decomposition of deformation energy, we obtain
total deformation energy, or generalized strain energy of the SDG,
1
U = λ̃ε j j εkk + μ̃ε jk ε jk + dγ[ jk] γ[ jk]
2
1
+ α 2 λ̃2 κ̂i j j κ̂ikk + α 2 μ̃2 κ̂i jk κ̂i jk + μ̃L2 χ jk χ jk , (45)
2
where
1 1
λ̃ = λ + (1 − α)2 d1 , μ̃ = μ + (1 − α)2 (d2 + d3 ) , d = (d2 − d3 ) . (46)
2 2
The first two terms in Eq. (45) are the strain energy density produced by macro-strain, which are the sum
of classical strain energy density and relative strain energy density. Due to the consistency of deformation
in experiment measurements, we define the sum of these two deformation energy densities as the total strain
energy density.
(ii) Definition of generalized stress and constitutive relations
Based on the generalized strain energy density in the SDG, we define generalized stresses, including Cauchy
stress τ jk , relative stress σ jk , superstress μi( jk) , and couple stress m jk ,
∂U ∂U
τ jk ≡ = τk j , σ̃ jk ≡ = −σ̃k j ,
∂ε jk ∂γ[ jk]
∂U ∂U
μ̃i( jk) ≡   = μ̃i(k j) , m jk ≡ = mk j . (47)
∂ α κ̂i jk ∂χ jk

By virtue of Eqs. (47) and (45), we obtain constitutive relations of the SDG,

τ jk = λ̃δ jk εll + 2μ̃ε jk , σ̃ jk = 2dγ[ jk] ,


μ̃i( jk) = α2 λ̃δ jk ∂i εll + 2μ̃∂i ε jk , m jk = 2μ̃L2 χ jk , (48)

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


2974 Y. Zhou, K. Huang

where λ̃ and μ̃ are nonlocal Lamé coefficients defined in the SDG.


(iii) Equilibrium equations and boundary conditions
The SDG based on the actual material assumption considers the micro-element subjected to forces and
moments, thus the equilibrium equations will not only be limited to the stress equilibrium equation, but
also the couple stress equilibrium equation. The variation of generalized strain energy density in the SDG is
 
δU = τ jk δε jk + σ[ jk] δω jk − δψ[ jk] + αμi( jk) ∂i δε jk + μi[ jk] ∂i δψ[ jk] . (49)

By integrating by parts, we have


 
δU = τ jk δε jk + σ[ jk] δω jk − δψ[ jk]
+ α∂i μi( jk) δε jk − α∂i μi( jk) δε jk
 
+ ∂i μi[ jk] δψ[ jk] − ∂i μi[ jk] δψ[ jk]
   
= ∂ j τ jk − α∂i μi( jk) + σ[ jk] δu k − ∂ j τ jk − α∂i μi( jk) + σ[ jk] δu k
   
+ ∂i μi jk δψ jk − σ[ jk] + ∂i μi[ jk] δψ[ jk] (50)

The volume integral of variation of generalized strain energy density over elastomer domain gives
 
   
δU = − ∂ j τ jk − α∂i μi( jk) + σ[ jk] δu k dV − σ[ jk] + ∂i μi[ jk] δψ[i j] dV
 
   
+ n j τ jk − α∂i μi( jk) + σ[ jk] δu k dS + n i μi jk δψ jk dS. (51)
∂ ∂

Based on Hamilton’s principle and by virtue of Eqs. (11) and (3), we obtain the stress equilibrium equation
and couple stress equilibrium equation,
 
∂ j τ jk − α∂i μi( jk) + σ[ jk] + f k = 0, ∂i μi[ jk] + σ[ jk] + Φ[ jk] = 0. (52)

Then stress boundary conditions and couple stress boundary conditions are derived,
 
tk = n j τ jk − α∂i μi( jk) + σ[ jk] , T jk = n i μi jk . (53)

On the other hand, we can equivalently give the couple stress equilibrium equations and boundary conditions
in terms of lower order tensors,

∂ j m ji + eikl σ[kl] + eikl Φ[kl] = 0, Tk = n j m jk . (54)



Define the total stress (or equilibrium stress) τ jk in the SDG, satisfying
◦  
τ jk = τ jk − α∂i μi( jk) = 1 − α 2 2 ∇ 2 τ jk + σ[kl] , (55)

For the small-scale deformation of some approximately uniform materials, that is, when we study some of
the size effect problem, we can reasonably ignore the relative deformation between the micro-elements and
the macro-body, and the nonlocal Cauchy stress is simplified into a symmetric form,
◦  
τ jk = τ jk − α∂i μi( jk) = 1 − α 2 2 ∇ 2 τ jk (56)

The equilibrium equations then simplify to


 
∂ j τ jk − α∂i μi( jk) + f k = 0, ∂i μi[ jk] + Φ[ jk] = 0. (57)

In this simplified form, the symmetric part μi( jk) and the anti-symmetric part μi[ jk] of the high-order stress are
decoupled. The solution to the problem is obtained by solving the stress equilibrium equation and the couple
stress equilibrium equation respectively, which greatly reduces the complexity of calculation.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


A simplified deformation gradient theory 2975

3.2 Degradations of the simplified deformation gradient theory

Under certain conditions, the SDG can be reduced to couple stress theory and strain gradient theory etc., when
the deformation of micro-elements is assumed more strictly. The degenerating process also clearly shows the
deformation effect of couple stress theory and strain gradient theory and gives clear physical significance.

3.2.1 Degenerate form I: the couple stress theory

In the SDG, assume that micro-elements rotate consistently with macro-body, i.e., ψ[ jk] = ω jk , the deforamtion
energy density associated with rotation degenerates to
(R) (R)
Ur(R) → U r = 0, Ug(R) → U g = a1 χi j χi j + a2 χi j χ ji + a3 χll2 . (58)

Also assume micro-elements are rigid and produce no strain, i.e., ψ( jk) = 0 and γ( jk) = ε jk , which corresponds
to the case of α = 0 in Eq. (31), to be specific,

(S) 1 1 (S)
Ur(S) → U r = d1 εii ε j j + (d2 + d3 ) εi j εi j , Ug(S) → U g = 0. (59)
2 2
The generalized strain energy density under this assumption is
1
U →U = λ̄εii ε j j + μ̄εi j εi j + a1 χi j χi j + a2 χi j χ ji + a3 χll2 , (60)
2
where λ̄ = λ + d1 , μ̄ = μ + (d2 + d3 ) /2. Ignoring the slight differences in the meaning of Lamé coefficients,
this is a classical form of couple stress theory.

3.2.2 Degenerate form II: the strain gradient theory

In the case where micro-elements deform consistently with macro-body, i.e., ψ( jk) = ε jk , then γ( jk) = 0,
which corresponds to the case of α = 1 in Eq. (31). The deformation energy density associated with strain
degenerates to

r(S) = 0, Ug(S) → U
g(S) = 1 2
Ur(S) → U λ̂ κ̂i j j κ̂ikk + μ̂2 κ̂i jk κ̂i jk . (61)
2
Considering micro-elements rotate synchronously with macro-body, i.e., ψ[ jk] = ω jk , γ[ jk] = 0, we have

r(R) = 0, Ug(R) → U
Ur(R) → U g(R) = a1 χi j χi j + a2 χi j χ ji + a3 χll2 . (62)

Then the generalized strain energy density becomes

 = 1 λ̂ε j j εkk + μ̂ε jk ε jk +


U →U
1 2
λ̂ κ̂i j j κ̂ikk + μ̂2 κ̂i jk κ̂i jk
2 2
+ a1 χ jk χ jk + a2 χ jk χk j + a3 χll2 , (63)

where λ̂ = λ, μ̂ = μ. The generalized strain energy density in Eq. (63) contains the deformation energy
generated by both strain gradient and rotation gradient, which is the strain energy density of Mindlin’s first
order strain gradient theory [21].
Eliminating the displacement from definitions in Table 2, we have κ̄i jk = κ̂ jki − κ̂ki j . The typical strain
gradient theory is a pure high-order deformation theory based on simple materials, which only includes the
gradient terms of macro-strain and does not contain terms related to rotation. Hence, the strain gradient theory
is still based on simple material models, and without considering the rotation of micro-elements. Only ε jk and
κ̂i jk left in the generalized strain energy density,

 = 1 λ̂ε j j εkk + μ̂ε jk ε jk + 1 λ̂2 κ̂i j j κ̂ikk + μ̂2 κ̂i jk κ̂i jk .


U →U (64)
2 2

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


2976 Y. Zhou, K. Huang

The new defined generalized stresses become


∂U ∂U 
τ̂ jk ≡ = τ̂k j , μ̂i jk ≡ = μ̂ik j , (65)
∂ε jk ∂ κ̂i jk
from which we obtain the simplified constitutive relations,

τ̂ jk = λ̂δ jk εll + 2μ̂ε jk , μ̂i jk = 2 λ̂δ jk ∂i εll + 2μ̂∂i ε jk = 2 ∂i τ̂ jk . (66)

The basic equations obtained under this degradation assumption are formally consistent with the strain gradient
theory, but the material parameters have slightly different meanings, which come from the relative deformation.

4 Material parameters and experimental verification of the SDG

In this section, a new explanation of elastic moduli is given based on the generalized strain energy, and the
nonlocal effect parameter is defined based on deformation energy to determine the nonlocal effect degree of
materials quantitatively. Nonlocal theoretical solutions to the tensile size effect of particle reinforced composites
and the micro-torsional size effect of cylinders are calculated based on the SDG and compared with experiment
results.

4.1 Material parameters and nonlocal effect parameters

4.1.1 New meaning and experimental comparison of material parameters

(i) New meaning of material parameters


The elastic moduli in classical local continuum theory are inherent properties of materials that is independent of
loadings and deformation forms. However, many experiments show that materials perform different mechanical
properties under different deformation and loading conditions. For example, in the tensile test of particle
reinforced composites, materials with different particle sizes exhibit different elastic moduli when the particle
volume fraction is constant.
In the deformation process of materials with nonlocal characteristics, the work of external loadings is not
only convert into classical strain energy, but also deformation energy related to the rotation of micro-elements.
For materials with different micro-structures under the same loading condition, the amount of work of external
loadings converted into strain energy is different, thus materials exhibit different elastic moduli. Therefore, the
interpretation of elastic moduli based on generalized strain energy can better reflect the mechanical properties
of actual materials.
According to the deformation hypotheses and corresponding deformation energy of the SDG in Sect. 3.1.1,
we rewrite the generalized strain energy density as

U = Us(T) + Ur(R) + Ug , (67)


(T) (S) (S)
where Us = Us + Ur is defined as the total strain energy density. Since Us and Ur are strain energies that
appear simultaneously in deformation and are both related to macro-strain, the classical Lamé coefficients and
parameters d1 − d3 defined in the SDG cannot be determined separately. Then we define λ̃ and μ̃ as nonlocal
Lamé coefficients in the SDG, which perform the same property as that in classical continuum theory.
We consider an object composed of actual materials, which deforms under the joint actions of external
forces and moments. Based on the classical continuum theory, deformation energy density stored in the object
is only classical strain energy density,
1
Uc = λc εii ε j j + μc εi j εi j . (68)
2
According to the SDG, strain energy, relative deformation energy and high-order gradient energy are generated
during the deformation of the object. For elastic deformation of the same object under same loadings, the elastic
deformation energy obtained by classical continuum theory and SDG must be consistent, i.e., conservation of
deformation energy, from which we can calibrate the material parameters in the SDG.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


A simplified deformation gradient theory 2977

For two independent deformations, the classical strain energy density are denoted as Uc , Uc and the gener-
alized strain energy of micro-structure system are U  , U  . According to conservation of deformation energy,
the classical strain energy calculated by using classical continuum theory must be equal to the generalized
strain energy on the basis of the SDG,

Uc = Us(T) + Ur(R) + Ug , Uc = Us(T) + Ur(R) + Ug , (69)


(T) (T)
in which, Us and Us are the total strain energy, satisfying
1  1
Us(T) = λ̃A + μ̃B  , Us(T) = λ̃A + μ̃B  , (70)
2 2
where
 

A = εii εj j dV, B =
εi j εi j dV,
 
A = εii εj j dV, B  = εij εij dV. (71)

Therefore, we obtain the relationship between the Lamé coefficients in the SDG and classical continuum
theory,
(R) (R)
2B  Ur + Ug − 2B  Ur + Ug
λ̃ = λc − ,
B  A − A B 
(R) (R)
−A Ur + Ug + A Ur + Ug
μ̃ = μc − . (72)
B  A − A B 
Due to the presence of rotation-related deformation energy in the micro-structure system, the total strain
energy in the SDG is always less than the strain energy in classical continuum theory, leading to the elastic
constants (Lamé coefficients in this paper) in the SDG are also smaller than that in the classical continuum
theory, as expressed by Eq. (72).
(ii) Experimental verification of Lamé coefficients in the SDG
The size effect experiment results of several particle reinforced metal matrix composites under uniaxial tension
[10–13] show that when the volume fraction of particles is determined, the strength of composites increases
with the decrease of particle size. The experiments also exhibit that the change of Young’s modulus with
particle sizes is obvious in the plastic deformation stage, but it can still be observed in the elastic deformation
stage. That is, the phenomenon of nonlocal effect still exists in the uniaxial tensile phenomenon in elastic
deformation, which is consistent with our predictions based on the micro-structure system.
In the uniaxial tensile test of particle reinforced composites, the properties of particles included in the
composite are similar to those of the micro-elements in the micro-structure system. We calibrate Lamé constants
in the SDG based on the experiment data of Yan et al. [13]. It must be noted here that we only study the elastic
phase of elastoplastic deformation in the experiment.
The materials used in the experiment are particle (SiC) reinforced metal (Al) matrix composites. The
percentage of SiC particles is 20%, and the size of SiC particles is 1 μm, 5 μm, 20 μm and 56 μm respectively.
The Young’s modulus and Poisson’s ratio of Al are E Al = 70 GPa and νAl = 0.33. The Young’s modulus and
Poisson’s ratio of SiC particles are E SiC = 450 GPa and νSiC = 0.17. For different composites with a certain
particle volume fraction, we think they are materials with the same component, and only particle sizes affect
the nonlocal effect of materials.
According to the hybrid model in composite material theory, the predicted Young’s modulus and Poisson’s
ratio of particle (SiC) reinforced metal (Al) matrix composites are 146 GPa and 0.3, respectively. The predicted
Young’s modulus based on the Halpin–Tsai formula [56] is E c = 101 GPa, which is closer to the actual value
of material. By virtue of the conversion between the elastic constants, we obtain corresponding predicted
classical Lamé coefficients, λc = 58.27 GPa and μc = 38.85 GPa.
Figure 4 gives the experimental true Lamé coefficients of SiC particles reinforced metal matrix composites
in elastic phase with particle sizes ranging from 1 μm to 56 μm under uniaxial tensile loadings. The Lamé
coefficients when particle size is close to diameter of an aluminum atom (λc , μc ) are the predicted values

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


2978 Y. Zhou, K. Huang

Fig. 4 The relation between true Lamé coefficients and particle sizes [13]

obtained by using the Halpin-Tsai formula. The values represented by the hollow points (λp , μp ) are predicted
values based on the SDG. The experiment results show that the nonlocal Lamé coefficients is approximately
linear to the particle sizes.
Figure 4 shows that the larger the particle sizes or element sizes, the stronger the nonlocality of materials. To
be specific, Lamé coefficients decrease as the particle sizes increase, or as the nonlocality of material increases.
Based on the SDG, when deformation of a body occurs, the micro-elements will produce relative deformation
that is inconsistent with the macro-deformation and dissipate part of total deformation energy, resulting in the
decrease of strain energy. Therefore, the corresponding Lamé coefficients are smaller than predicted classical
Lamé coefficients. On account of conservation of deformation energy, the difference between strain energy in
the SDG and that in classical continuum theory is interpreted as the nonlocal deformation energy.
Rewrite the relative rotation energy density in terms of the average relative rotation components of particles
(R)
under the statistical idea, Ur = d γ [i j] γ [i j] . For actual materials, we cannot accurately measure the rotation
deformation of micro-elements in specific experiments, and we take its average instead. Assume that the
average relative rotation components γ [i j] is proportional to one half power of particle size D, i.e.,

γ [i j] = di j D, (73)

whence

Ur(R) = ddi2j D ≡ dT D. (74)

Equation (74) indicates that the relative rotation energy density is proportional to the first power of particle
size D.
In uniaxial tensile deformation, the macro-strain components are constants that are independent of position
coordinates. The macro-rotation components are equal to zero. Then, we have

κ̂i jk = ∂i ε jk = 0, κ̄i jk = ∂i ω jk = 0, (75)


(S) (R)
thus Ug = Ug = 0. According to the conservation of deformation energy, we obtain

Us(T) + Ur(R) = Uc , (76)

where Uc is the strain energy density converted from all external work in classical continuum theory.
In uniaxial tensile experiment measuring the elastic moduli of material in x1 direction, the difference
between strain energy density in the classical continuum theory and total strain energy density in the SDG
gives relative rotation energy density since the high-order gradient energies are vanished,
   
(R) 1 (1 − ν) E c − E p 2
Ur = λc − λ̃ + (μc − μ̃) ε11 =
2
ε . (77)
2 2(1 + ν)(1 − 2ν) 11
By virtue of Eqs. (74) and (77), Young’s modulus or Lamé coefficients should decrease linearly with particle
size.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


A simplified deformation gradient theory 2979

The experiment of Yan et al. [13] proves the existence of nonlocal rotation effect and shows that the
relationship between Young’s modulus and particle sizes during uniaxial tension is approximately linear.
Through linear regression fitting, we get
E p (D) = −0.1518 D + 80.15, (78)
where the dimensions of predicted Young’s modulus and particle size are GPa and μm. The correlation
coefficient of linear regression is R 2 = 0.9995, which shows that Eq. (78) well reflects the linear relationship
between Young’s modulus and particle sizes on the micron scale.
Remark (i) In the simple tensile deformation, the macro-strain components are constants, and the macro-
rotation components are zero. Thus, the high-order gradient energies in strain gradient theory or couple stress
theory are vanished. The strain energy density given by these two theories is the same as that by classical
continuum theory, and the size effect cannot be explained, while the relative rotation energy introduced in the
SDG can well explain the size effect of elastic moduli at different particle sizes.
(ii) Based on the reasonable assumptions of micro-deformation, relative deformation is related to macro-
strain, which facilitates the calculation of nonlocal deformation energy. Since part of the relative deformation
energy is coupled to the classical strain energy, we cannot accurately quantitatively predict the value of elastic
constants, but Lamé coefficients defined in SDG are closer to the true value because the nonlocal deformation
energy is taken into account.
(iii) In the SDG model, there are other forms of deformation energy besides the classical strain energy,
and the prediction of Lamé coefficients is based on conservation of deformation energy. Therefore, Lamé
coefficients are related to types of external loading and forms of deformation, and cannot be simply understood
as the basic constant parameters of materials.

4.1.2 Definition and experimental determination of nonlocal effect parameter

Decompose the high-order gradient energy density into strain gradient energy density Ug(S) and rotation gradient
(R)
energy density Ug , and rewrite the generalized strain energy density as

U = Us(T) + Ur(R) + Ug(S) + Ug(R) , (79)


and then we may give the definition of dimensionless nonlocal effect parameter N ,
N ≡ Un /U, (80)
where Un is total nonlocal deformation energy density, satisfying

Un = Ur(R) + Ug(S) + Ug(R) . (81)


We can define the nonlocal effect parameters more specifically corresponding to each partial nonlocal energy
as
N = Nr(R) + Ng(S) + Ng(R) = Ur(R) /U + Ug(S) /U + Ug(R) /U, (82)

where Nr(R) , Ng(S) and Ng(R) are nonlocal relative rotation effect parameter, high-order strain gradient effect
parameter and high-order rotation gradient effect parameter respectively.
On account of Eqs. (76) and (78), we have

0.1518(1 − ν)ε11
2
Ur(R) = D. (83)
2(1 + ν)(1 − 2ν)

Let ε11 = 0.8 × 10−3 , the predicted classical strain energy density is Uc = 3.433 × 10−5 GPa. Equation
(76) indicates that the difference between total strain energy in SDG and classical strain energy is the relative
rotation energy in uniaxial tensile deformation. Based on the elastic phase of Yan’s experiment [13], Table 3
gives the variation trend of strain energy density and relative rotation energy density with different particle
sizes under the same composition in uniaxial tensile test. With the increase of particle sizes, the total strain
energies gradually decrease, while the relative rotation energies increase, i.e., the nonlocal effects become

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


2980 Y. Zhou, K. Huang

Table 3 Nonlocal deformation energy densities and nonlocal effect parameters in uniaxial tensile deformation [13]

Energy Uniform size D = 1 μm D = 5 μm D = 20 μm D = 56 μm


(T)
Us (GPa) 3.433 × 10−5 3.428 × 10−5 3.403 × 10−5 3.298 × 10−5 3.071 × 10−5
(R)
Ur (GPa) 0 0.005 × 10−5 0.030 × 10−5 0.136 × 10−5 0.362 × 10−5
(R)
N Nr 0 0.001 0.009 0.040 0.106

stronger with the increase of particle sizes, or the nonlocal material moduli become stronger with the decrease
of particle sizes.
According to Table 3, the dimensionless nonlocal effect parameters increase with the increase of particle
sizes, which shows that the nonlocal effect of material is enhanced. To be specific, when the particle size
is 56 μm, the material exhibits relatively strong nonlocal effect, and the nonlocal rotation effect parameter
reaches about 0.1.
Remark (i) The materials perform stronger nonlocal effects when the particle size becomes greater. Therefore,
in the mechanical research of some materials, we cannot simply ignore the nonlocal effect energy, which
cannot reflect the true properties of material.
(ii) Nonlocal effect parameter is defined as the ratio of nonlocal energy to generalized strain energy, which
is used to measure the proportion of energy consumed by nonlocal deformation to total deformation energy.
The nonlocal effect parameter of ideal uniform continuum material is zero, and the larger nonlocal effect
parameter is, the stronger nonlocality of material is, resulting in the reduction of material moduli.

4.2 Theoretical solution and experimental verification of nonlocal torsion

Experiments have shown that the micro-torsion of cylinders performs obvious size effect, i.e., the torsion
stiffness increases significantly with the decrease of diameter. Many scholars have studied the size effect of
torsion through strain gradient theory, couple stress theory or other nonlocal theories. The theoretical results
can explain the size effect phenomenon qualitatively, but the quantitative values are often slightly different
from the experiment results. For the micro-deformation of uniform materials, based on the experiment results
in Sect. 4.1, the nonlocal relative rotation effect is weak, and its influence can be reasonably ignored in the study
of micro-torsion. In this section, the SDG is applied to theoretical calculation and experimental comparison
of size effect of nonlocal torsion of cylinders, and the theoretical predictions of torsion stiffness under smaller
diameters are given.

4.2.1 Torsion stiffness of a cylinder

In the Cartesian coordinate system (x1 , x2 , x3 ), we study relationship between nonlocal torsion stiffness and
diameters of cylinders. Let the axial of cylinder be along x3 direction, and assume that displacement field in a
cylinder is the same as that in classical continuum theory,
u 1 = −ϑ x2 x3 , u 2 = ϑ x1 x3 , u 3 = ϑϕ(x1 , x2 ), (84)
where ϑ is the torsion angle per unit length of cylinders, and ϕ(x1 , x2 ) is the St. Venant torsion function,
satisfying the Laplace equation, i.e., ∇ 2 ϕ(x1 , x2 ) = 0. Based on Eq. (84), the components of macro-strain
tensor are
   
ϑ ∂ϕ ϑ ∂ϕ
ε23 = + x1 , ε31 = − x2 , ε11 = ε22 = ε33 = ε12 = 0. (85)
2 ∂ x2 2 ∂ x1
According to Table 2, non-zero components of strain gradient tensor can be obtained,
 2 
ϑ ∂ ϕ ϑ ∂ 2ϕ
κ̂123 = + 1 , κ̂223 = ,
2 ∂ x1 ∂ x2 2 ∂ x22
 2 
ϑ ∂ ϕ ϑ ∂ 2ϕ
κ̂231 = − 1 , κ̂131 = . (86)
2 ∂ x1 ∂ x2 2 ∂ x12

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


A simplified deformation gradient theory 2981

Fig. 5 Relationship between dimensionless torsion stiffness and micro-strain proportionality coefficient and dimensionless radius

By virtue of Eq. (84), components of macro-rotation tensor are


   
ϑ ∂ϕ ϑ ∂ϕ
w1 = − x 1 , w2 = − − x 2 , w3 = ϑ x 3 , (87)
2 ∂ x2 2 ∂ x1
and non-zero components of symmetric curvature tensor can be then obtained,
 2   
ϑ ∂ ϕ ϑ ∂ 2ϕ
χ11 = − 1 , χ22 = − −1 ,
2 ∂ x1 ∂ x2 2 ∂ x1 ∂ x2
 
ϑ ∂ ϕ 2 ∂ ϕ
2
χ33 = ϑ, χ12 = − 2 . (88)
4 ∂ x2 2 ∂ x1

In the case where the cross section of cylinders is circular, the St. Venant torsion function ϕ(x1 , x2 ) = 0.
Applying Eqs. (38) and (42), we have the strain gradient energy density and rotation gradient energy density
for nonlocal torsion deformation,

Ug(S) = α 2 μ̃2 ϑ 2 , Ug(R) = 3μ̃L2 ϑ 2 . (89)

On account of the relationship between external force work and generalized strain energy, we obtain torsion
stiffness of cylinders based on the SDG,

T = T0 1 + 2α 2 R̂ −2 + 6 R̄ −2 , (90)

where R̂ = R/ and R̄ = R/L are dimensionless strain radius and dimensionless rotation radius respectively,
R is the radius of cylinders, and T0 = π μ̃ϑ R 4 /2 is the torque per unit length of cylinders in classical continuum
theory. The middle term and last term in Eq. (90) represent the effect of strain gradient and rotation gradient
on torsion stiffness respectively. The analytical solution to nonlocal torsion based on the SDG shows that both
strain gradient and rotation gradient have effects on the torsion stiffness of cylinders, and the rotation gradient
has a greater effect than strain gradient.

4.2.2 Nonlocal effects of torsion stiffness

In the SDG, the micro-strain proportionality coefficient α has an important influence on the deformation of
micro-element and magnitude of generalized strain energy. The high-order gradient energy in generalized strain
energy reduces to the high-order deformation energy in couple stress theory when α = 0 and corresponds to
the high-order deformation energy in Mindlin’s first-order strain gradient theory when α = 1. For torsion of
fine copper wire, the length-scale parameters  and L both take 3 μm [57]. Let R̂ = R̄ = R0 , and R0 denotes
the dimensionless radius of fine copper wire.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


2982 Y. Zhou, K. Huang

Table 4 Experimental comparison of size effect on torsion stiffness of fine copper wire

2R/μm 170 30 20 15 12
R0 0.035 0.2 0.3 0.4 0.5
T /T0 1.00997 1.32 1.72 2.28 3.0

Table 5 Theoretical prediction of size effect on torsion stiffness of fine copper wire

2R/μm 10 8 6
R0 0.6 0.75 1.0
T /T0 3.88 5.5 9.0

Figure 5 exhibits the relationship between dimensionless torsion stiffness and dimensionless radius R0
when α = 0, 0.7, 1.0, respectively. Under the case where dimensionless radius R0 = 1.0, i.e. the radius of fine
copper wire is equal to the characteristic scale, the dimensionless T /T0 are 7.0 and 9.0 when α are equal to 0
and 1 respectively, where the difference is contribution of strain gradient to torsion stiffness. The difference
between torsion stiffness when α = 0 and classical continuum theory only comes from the contribution of
rotation gradient. The effect of α on the torsion stiffness decreases with the increase of dimensionless radius
R0 , and when R0 is smaller than 0.1, the effect of strain gradient and rotation gradient on the torsion stiffness
tends to vanish gradually.
According to Eq. (90), the elastic torsion deformation of fine copper wire is affected by both strain gradient
and rotation gradient at micron scale. When the radius of copper wire is equivalent to the characteristic scale,
the torsion stiffness increases by 2.0 times due to strain gradient, and 6.0 times due to rotation gradient, and
the total nonlocal torsion stiffness is about 9.0 times of that without considering size effect.
One of the most famous experiment of micro-torsion is the fine copper wire tested by Fleck et al. [58]. The
diameters of copper wire reduce from 170 to 12 μm, and the torsion stiffness increases by 3.0 times, which is
in good agreement with theoretical solutions based on the SDG in Table 4.
The experiment of Fleck et al. [58] shows that the size effect of micro-torsion is obvious when R0 is greater
than 0.2, while when R0 is relatively small, the micro-torsion stiffness almost does not perform size effect,
which is consistent with theoretical results in Table 4. Furthermore, we predict the nonlocal torsion stiffness
of copper wire with smaller diameters (0.6  R0  1.0), which is given in Table 5.

5 Conclusions

We have presented a simplified nonlocal deformation gradient theory with two length-scale parameters and
explained the deformation and energy effects of nonlocal theories in detail. In the SDG, materials are con-
sidered as a continuum model composed of micro-elements, and the generalized strain energy is introduced
to describe the total deformation energy of micro-structure system. The generalized strain energy includes
relative deformation energy and high-order gradient energy besides classical strain energy. By virtue of the
generalized strain energy, elastic moduli of materials are reinterpreted from the perspective of deformation
energy, and the nonlocal effect parameter is defined to quantitatively measure the degree of nonlocal effect of
materials. As a simplified nonlocal theory, the SDG can also be degenerated into couple stress theory, strain
gradient theory and classical elastic theory.
The SDG performs well in solving two types of nonlocal deformation problems, including the macro-
deformation of non-uniform materials and the micro-deformation of materials with micro-structures. For the
macroscopic uniaxial tension of particle reinforced composites, the SDG predicts the equivalent elastic moduli
and approximatively linear relationship on the micron scale between Lamé coefficients and particle sizes under
the condition of a certain composition, which is verified by experiments. In micro-torsion of fine copper wire,
the nonlocal torsion stiffness of diameter 12 μm obtained by the SDG is about 3.0 times that of without
considering size effect, which is in good agreement with the experiment results. We also predict nonlocal
torsion stiffness at smaller diameters (6 μm  2R  10 μm). The results show that when the radius of copper
wire is equal to the characteristic scale, the nonlocal torsion stiffness is 9.0 times of classical value.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


A simplified deformation gradient theory 2983

Funding Funding is provided by National Social Science Fund of China (Grant No. 11521202).

References

1. Tang, C.Z., Alici, G.: Evaluation of length-scale effects for mechanical behaviour of micro-and nanocantilevers: I. Experi-
mental determination of length-scale factors. J. Phys. D. Appl. Phys. 44(33), 335501 (2011)
2. Tang, C.Z., Alici, G.: Evaluation of length-scale effects for mechanical behaviour of micro- and nanocantilevers: Ii. Experi-
mental verification of deflection models using atomic force microscopy. J. Phys. D. Appl. Phys. 44(33), 335502 (2011)
3. Lei, J., He, Y., Guo, S., Li, Z., Liu, D.: Size-dependent vibration of nickel cantilever microbeams: experiment and gradient
elasticity. Aip Adv. 6(10), 51–59 (2016)
4. Li, Z.K., He, Y.M., Lei, J., Guo, S., Liu, D.B., Wang, L.: A standard experimental method for determining the material length
scale based on modified couple stress theory. Int. J. Mech. Sci. 141, 198–205 (2018)
5. Stolken, J.S., Evans, A.G.: A microbend test method for measuring the plasticity length scale. Acta Mater. 46(14), 5109–5115
(1998)
6. Anderson, W.B., Lakes, R.S.: Size effects due to cosserat elasticity and surface damage in closed-cell polymethacrylimide
foam. J. Mater. Sci. 29(24), 6413–6419 (1994)
7. Andrews, E.W., Gioux, G., Onck, P., Gibson, L.J.: Size effects in ductile cellular solids. Part ii: experimental results. Int. J.
Mech. Sci. 43(3), 701–713 (2001)
8. Chong, A.C., Lam, D.C.: Strain gradient plasticity effect in indentation hardness of polymers. J. Mater. Res. 14(10), 4103–
4110 (1999)
9. Bastawros, A.F., Bart, S.H., Evans, A.G.: Experimental analysis of deformation mechanisms in a closed-cell aluminum alloy
foam. J. Mech. Phys. Solids 48(2), 301–322 (2000)
10. Yang, J., Cady, C., Hu, M.S., Zok, F., Mehrabian, R., Evans, A.G.: Effects of damage on the flow strength and ductility of a
ductile al alloy reinforced with sic particulates. Acta Metal. Et. Mater. 38(12), 2613–2619 (1990)
11. Lloyd, D.J.: Particle-reinforced aluminum and magnesium matrix composites. Int. Mater. Rev. 39(1), 1–23 (1994)
12. Kouzeli, M., Mortensen, A.: Size dependent strengthening in particle reinforced aluminium. Acta Mater. 50(1), 39–51 (2002)
13. Yan, Y.W., Geng, L., Li, A.B.: Experimental and numerical studies of the effect of particle size on the deformation behavior
of the metal matrix composites. Mater. Sci. Eng. A 448(1–2), 315–325 (2007)
14. Eringen, A.C., Edelen, D.: On nonlocal elasticity. Int. J. Eng. Sci. 10(3), 233–248 (1972)
15. Krner, E.: Elasticity theory of materials with long range cohesive forces. Int. J. Solids Struct. 3(5), 731–742 (1967)
16. Kunin, I.A.: Elastic Media with Microstructure. Springer, Berlin (1983)
17. Silling, S.A.: Reformulation of elasticity theory for discontinuities and long-range forces. J. Mech. Phys. Solids 48(1),
175–209 (2000)
18. Rogula, D.: Nonlocal Theory of Material Media. Springer, Berlin (1982)
19. Kunin, I.A.: Elastic Media with Microstructure I: One-Dimensional Models. Springer Berlin Heidelberg (1982) https://doi.
org/10.1007/978-3-642-81748-9
20. Mindlin, R.D.: Second gradient of strain and surface-tension in linear elasticity. Int. J. Solids Struct. 1(4), 417–438 (1965)
21. Mindlin, R.D., Eshel, N.N.: On first strain-gradient theories in linear elasticity. Int. J. Solids Struct. 4(1), 109–124 (1968)
22. Eringen, A.C.: On differential equations of nonlocal elasticity and solutions of screw dislocation and surface waves. J. Appl.
Mech. 54(9), 4703–4710 (1983)
23. Aifantis, E.C.: On the role of gradients in the localization of deformation and fracture. Int. J. Eng. Sci. 30(10), 1279–1299
(1992)
24. Lazar, M., Maugin, G.A.: Nonsingular stress and strain fields of dislocations and disclinations in first strain gradient elasticity.
Int. J. Eng. Sci. 43(13/14), 1157–1184 (2005)
25. Cosserat, E.: Theorie des corpes deformables. Nature 81, 67 (1909)
26. Voigt, W.: Theoretische studien uber die elasticitatsverhaltnisse der krystalle. Annalen der Physik, 38, 573–587 (1889)
27. Toupin, R.A.: Elastic materials with couple-stresses. Arch. Ration. Mech. An. 11(1), 385–414 (1962)
28. Koiter, W.T.: Couple-stress in the theory of elasticity. Int. J. Solids Struct. 67, 17–44 (1963)
29. Mindlin, R.D., Eshel, N.N.: Effects of couple-stresses in linear elasticity. Int. J. Solids Struct. 1, 417–438 (1968)
30. Yang, F., Chong, A., Lam, D., Tong, P.: Couple stress based strain gradient theory for elasticity. Int. J. Solids Struct. 39(10),
2731–2743 (2002)
31. Park, S.K., Gao, X.L.: Bernoulli–Euler beam model based on a modified couple stress theory. J. Micromech. Microeng.
16(11), 2355–2359 (2006)
32. Shaat, M.: Physical and mathematical representations of couple stress effects on micro/nanosolids. Int. J. Appl. Mech. 07(01),
1550012 (2015)
33. Neff, P., Munch, I., Ghiba, I.D., Madeo, A.: On some fundamental misunderstandings in the indeterminate couple stress
model. A comment on recent papers of a.r hadjesfandiari and g.f. dargush. Int. J. Solids Struct. 81, 233–243 (2016)
34. Tsiatas, G.C.: A new Kirchhoff plate model based on a modified couple stress theory. Int. J. Solids Struct. 46(13), 2757–2764
(2009)
35. Chen, W.J., Li, X.P.: A new modified couple stress theory for anisotropic elasticity and microscale laminated Kirchhoff plate
model. Arch. Appl. Mech. 84(3), 323–341 (2014)
36. Deng, G.Q., Dargush, G.: Mixed convolved lagrange multiplier variational formulation for size-dependent elastodynamic
couple stress response. Acta Mech. 233(5), 1837–1863 (2022)
37. Wang, Y.X., Zhang, X., Shen, H.M., Liu, J., Zhang, B.: Couple stress-based 3d contact of elastic films. Int. J. Solids Struct.
191–192, 449–463 (2020)
38. Liu, N., Fu, L.Y., Tang, G., Kong, Y., Xu, X.Y.: Modified lsm for size-dependent wave propagation: comparison with modified
couple stress theory. Acta Mech. 231(4), 1285–1304 (2020)

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


2984 Y. Zhou, K. Huang

39. Apostolakis, G., Dargush, G.F.: Size-dependent couple stress natural frequency analysis via a displacement-based variational
method for two-and three-dimensional problems. Acta Mech. 234, 891–910 (2022)
40. Aifantis, E.C.: On the microstructural origin of certain inelastic models. J. Eng. Mater. Technol. 106(4), 326–330 (1984)
41. Fleck, N.A., Hutchinson, J.W.: Strain gradient plasticity. Adv. Appl. Mech. 33, 295–361 (1997)
42. Han, C.S., Gao, H.J., Huang, Y.G., Nix, W.D.: Mechanism-based strain gradient plasticity-i. Theory. J. Mech. Phys. Solids
47(5), 1239–1263 (1999)
43. Aifantis, E.C.: Update on a class of gradient theories. Mech. Mater. 35(3/6), 259–280 (2003)
44. Fleck, N.A., Hutchinson, J.W.: A reformulation of strain gradient plasticity. J. Mech. Phys. Solids 49(10), 2245–2271 (2001)
45. Lam, D., Yang, F., Chong, A., Wang, J., Tong, P.: Experiments and theory in strain gradient elasticity. J. Mech. Phys. Solids
51(8), 1477–1508 (2003)
46. Zhou, S.J., Li, A.Q., Wang, B.L.: A reformulation of constitutive relations in the strain gradient elasticity theory for isotropic
materials. Int. J. Solids Struct. 80, 28–37 (2016)
47. Fu, G.Y., Zhou, S.J., Qi, L.: On the strain gradient elasticity theory for isotropic materials. Int. J. Eng. Sci. 154, 103348
(2020)
48. Rahimi, Z., Rezazadeh, G., Sumelka, W.: A non-local fractional stress-strain gradient theory. Int. J. Mech. Mater. Des. 16(2),
265–278 (2020)
49. Fu, G.Y., Zhou, S.J., Qi, L.: A size-dependent Bernoulli–Euler beam model based on strain gradient elasticity theory incor-
porating surface effects. ZAMM 99(6), 1–22 (2019)
50. Jiang, Y.Y., Li, L., Hu, Y.J.: Strain gradient elasticity theory of polymer networks. Acta Mech. 233(8), 3213–3231 (2022)
51. Barretta, R., Faghidian, S.A., Marotti de Sciarra, F.: Aifantis versus lam strain gradient models of bishop elastic rods. Acta
Mech. 230(8), 2799–2812 (2019)
52. Lazar, M.: Incompatible strain gradient elasticity of mindlin type: screw and edge dislocations. Acta Mech. 232(9), 3471–3494
(2021)
53. Le, T.M., Vo, D., Rungamornrat, J., Bui, T.Q.: Strain-gradient theory for shear deformation free-form microshells: governing
equations of motion and general boundary conditions. Int. J. Solids Struct. 248, 111579 (2022)
54. Li, G.E., Kuo, H.Y.: Effects of strain gradient and electromagnetic field gradient on potential and field distributions of
multiferroic fibrous composites. Acta Mech. 232(4), 1353–1378 (2021)
55. Mindlin, R.D.: Micro-structure in linear elasticity. Arch. Ration Mech. An. 16(1), 51–78 (1964)
56. Ashton, J.E., Halpin, J.C., Petit, A.: Primer on Composite Materials: Analysis. Washington University, Washington (1969)
57. Chong, A.C., Yang, F., Lam, D.C.: Torsion and bending of micron-scaled structures. J. Mater. Res. 16(4), 1052–1058 (2001)
58. Fleck, N.A., Muller, G.M., Ashby, M.F., Hutchinson, J.W.: Strain gradient plasticity: theory and experiment. Acta Metal
Mater. 42(2), 475–487 (1994)

Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional
affiliations.

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement
with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely
governed by the terms of such publishing agreement and applicable law.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:

1. use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
2. use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
3. falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
4. use bots or other automated methods to access the content or redirect messages
5. override any security feature or exclusionary protocol; or
6. share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at

onlineservice@springernature.com

You might also like