Next Article in Journal
Force and Pressure Dependent Asymmetric Workspace Research of a Collaborative Robot and Human
Previous Article in Journal
An Overview of the Searches for the Violation of the Charge-Parity Symmetry in the Leptonic Sector
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Explicit Formulas for the Deformation of Chiral Porous Circular Beams in Gradient Thermoelasticity

Department of Structures for Engineering and Architecture, University of Naples Federico II, Via Forno Vecchio, 80131 Naples, Italy
Symmetry 2024, 16(1), 129; https://doi.org/10.3390/sym16010129
Submission received: 2 January 2024 / Revised: 16 January 2024 / Accepted: 18 January 2024 / Published: 22 January 2024
(This article belongs to the Section Engineering and Materials)

Abstract

:
Chirality and porosity are characteristic properties of nanostructured materials. Their effects on the mechanical behaviour of structural elements, such as shells, plates and beams, cannot be disregarded. In this paper, we study the thermoelastic deformation of a chiral porous circular beam loaded with an axial force and torque. The beam is also under the action of a constant temperature field. The analytical solution is obtained using the results established in a paper recently published by the Author within the context of the strain gradient theory proposed by Papanicopolous. In the constitutive equations, the chirality is introduced by a material constant parameter and the porosity is described by means of a scalar function. Displacements, microdilatation function, and stress and strain fields are expressed in explicit form and in terms of engineering constants. Explicit formulas of the stiffness of chiral porous circular beams are presented and the effects of right and left chirality are discussed.

1. Introduction

In general, the term chiral, is used to describe an object that is non-superposable on its mirror image. In continuum mechanics, a material whose properties are not invariant to inversion is called chiral. Ordinary materials, as structural and building materials, do not exhibit chirality and this property has been ignored in mechanics for a long time. In recent years, the great scientific and technical achievements in the creation of nanomaterials and nanotechnology have driven interest in the research of novel phenomena and properties of materials whose internal structure has nanoscale dimensions. Chirality is a characteristic property of nanomaterials and its effects on the behaviour of structural elements (shells, plates, beams, etc.) are not negligible. For a brief historical sketch on nanomaterials and nanotechnology, see Guz and Rushchitskii [1] and the references therein. For an overview of chirality in mechanics, see Lakes [2].
Various mathematical theories have been considered to introduce chirality in the mechanical behaviour of materials [3,4,5,6]. Among the microcontinuum theories of elasticity, the Cosserat theory has attracted increasing interest for describing microstructural effects due to chirality of the microstructure [7,8,9]. An extended micropolar theory was ideated to study the chiral effects that arise from coupling between stretching deformations and the microrotation [10,11,12,13]. In Cosserat models, chirality introduces three additional parameters for a total of nine material parameters. An alternative approach is provided by the strain gradient theory, which has been shown to be effective in investigating problems related to size effects and nanotechnology [14,15]. In [16], Papanicolopulos adapted the Mindlin–Eshel strain gradient theory for centrosymmetric materials [17] for the case of noncentrosymmetric solids. In contrast with the Cosserat model, in the strain gradient theory the chiral behaviour is controlled by a single material parameter and the sign of parameter allows for distinguishing between right and left chirality.
Porosity is another important feature of materials with nanostructures. Nanotubes, nanospheres, and other nanoparticles possess a high porosity because of their internal cavities [18,19].
There are a number of theories that describe the mechanical properties of materials with single and multiple porosity (see, e.g., [20,21,22] and the references therein). Microcontinuum field theories provide suitable mathematical models to investigate the behaviour of porous media [23,24]. For instance, if in the microstretch theory the rigid microrotation of the material particles is absent, we obtain the theory of materials with voids proposed by Cowin and Nunziato [25]. In both theories, the variation in void volume is described by an independent kinematical variable named the microdilatation function. The Cowin–Nunziato theory has been subject to intensive study and a great number of contributions regarding the fundamentals and applications have been published [26,27,28,29,30,31,32,33].
This paper is concerned with the strain gradient theory of porous thermoelastic materials. The theory is constructed adding the second-order partial derivative of the displacement and the first order derivative of the microdilatation function in the classical set of independent constitutive variables [34,35]. We applied this theory to study the thermoelastic deformation of a chiral porous circular cylinder.
The problems of the elastic deformation of generalized models of beams have been largely investigated [36,37,38,39,40,41]. For recent theoretical results on the theory of nanobeams, see [42,43,44,45]. The method of solution is well known and consists of reducing the 3D problem in the solution of two-dimensional problems. Usually, the solution depends on the cross section and cannot be expressed in explicit form, so we have only a qualitative description of the deformation. In the case of chiral beams, in contrast with the classical elasticity, an axial force produces torsion and torque produces extension. These effects are not typical of chirality and have been observed in some nonlinear phenomena, termed the poynting effect [46,47,48]. Possible correlations between the poynting effect and chirality have not yet been investigated.
In a recent article, De Cicco and Iesan [49] solved the general problem of the thermoelastic deformation of chiral porous cylinders. In this paper, we used the results established in [49] to obtain explicit formulas for the case of a cylinder with a circular cross-section and loaded with an axial force, torque, and constant temperature field. As the general solution presented in [49] was quite complex and could be difficult to use, the purpose of this study was to express the relations for the displacement, microdilatation function, and stress and strain fields in a convenient form, avoiding the use exceedingly complicated and tedious calculations used by researchers in the field. We obtained a clear qualitative and quantitative description of the deformation, which made it possible to distinguish the effects of right and left chirality. Using suitable notations, we introduced engineering constants and expressed the solution in terms of the Young-type modulus and Poisson-type ratio. The results were of practical use and would be useful for testing mechanical properties, such as the stiffness and strength of materials with nanostructures.
The paper is structured as follows. First, we present the basic equations of chiral porous elastic solids and formulate the problem of thermoelastic deformation of the right circular cylinder. Then, we establish the solution with the help of two-dimensional problems. The solutions of special cases, such as chiral (non porous) cylinder and porous (non chiral) cylinder, are derived. Using appropriate notations the displacements, microdilatation function, and stress and strain fields are expressed in terms of engineering constants. The results and the effects of right and left chirality are discussed.

2. Preliminaries

In this section, we formulate the equilibrium problem of a porous chiral cylinder subjected to mechanical forces and a variation in temperature. The cylinder is supposed to be homogeneous and isotropic. We denote the lateral boundary by Π , the terminal cross-sections by Σ α ( α = 1 , 2 ) , a generic cross-section by Σ , the boundary of Σ α by Γ α , the boundary of Σ by Γ , and the length of the cylinders by h. We chose a system of rectangular axes, such that x 3 a x i s is parallel to the generator of the cylinder and the x 1 O x 2 p l a n e contains the basis Σ 1 at x 3 = 0 . Let F = ( F 1 , F 2 , F 3 ) and M = ( M 1 , M 2 , M 3 ) be the prescribed vector representing the resultant force and the resultant moment about O of the tractions acting on Σ 1 . On Σ 2 , there are tractions applied so as to satisfy the equilibrium conditions of the cylinder. We assume a thermal field T that is independent of the axial coordinate x 3 . The cylinder is supposed to be free of lateral loading. Let u i be the components of the displacement vector and φ is the microdilatation function. The equilibrium problem of the cylinder consists of finding the functions u i and φ , satisfying the following systems of equations:
geometric equations
e i j = 1 2 ( u i , j + u j , i ) , κ i j k = u k , i j
where e i j is the strain tensor and κ i j k is the strain gradient tensor.
constitutive equations
τ i j = λ e r r δ i j + 2 μ e i j + d φ δ i j + f ( ε i k m κ j k m + ε j k m κ i k m ) b T δ i j , μ i j k = 1 2 α 1 ( κ r r i δ j k + 2 κ k r r δ i j + κ r r j δ i k ) + α 2 ( κ i r r δ j k + κ j r r δ i k ) + 2 α 3 κ r r k δ i j + β 1 δ i j φ , k + β 2 ( δ i k φ , j + δ j k φ , i ) + 2 α 4 κ i j k + α 5 ( κ k j i + κ k i j ) + f ( ε i k s e j s + ε j k s e i s ) , σ i = β 1 κ r r i + 2 β 2 κ i r r + a 0 φ , i , g = d e r r + ξ φ β T ,
where τ i j is the stress tensor; μ i j k is the dipolar stress tensor; σ i is the equilibrated stress vector; g is the intrinsic body force; T is the temperature; δ i j is the Kronecker delta; ε i j k is the alternating symbol; λ , μ and b are the constitutive constants of the classical theory of elasticity; α i ( i = 1 , 2 , , 5 ) and β j ( j = 1 , 2 ) are the constitutive constants associated with the gradient terms; d , a 0 , ξ , and β are constitutive constants linked to porosity; and f is a constant associated with the chiral behaviour.
equilibrium equations
τ j i , j μ k j i , k j = 0 , σ j , j g = 0 .
The following boundary conditions
P i = 0 , R i = 0 , Q i = 0 , σ α n α = 0
must be satisfied on the lateral surface Π . Here, ( n 1 , n 2 , 0 ) are the direction cosines of the exterior normal to Π and
P i = τ 3 i + 2 μ α 3 i , α + μ 33 i , 3 , R i = μ 33 i on Σ 1 , Q i = 2 μ α 3 i n α , on Γ 1 .
Further,
Σ 1 P α d a + Γ 1 Q α d s = F α ,    
Σ 1 P 3 d a + Γ 1 Q 3 d s = F 3 ,    
Σ 1 ( x α P 3 + R α ) d a + Γ 1 x α Q 3 d s = ε β α 3 M β ,
Σ 1 ε α β 3 x α P β d a + Γ 1 ε α β 3 x α Q β d s = M 3      
have to be satisfied on the end Σ 1 .

3. Circular Cylinder

We consider a porous chiral circular cylinder of radius a subjected to the following external data
F = ( 0 , 0 , F 3 ) , M = ( 0 , 0 , M 3 ) , T = T *
where F 3 , M 3 , and T * are the prescribed constants. The solutions of the problem under discussion have the form
u α = ε 3 β α a 2 x β x 3 + a 1 u α ( 1 ) + a 2 u α ( 2 ) + w α , u 3 = a 1 x 3 + a 1 u 3 ( 1 ) + a 2 u 3 ( 2 ) + w 3 , φ = a 1 φ ( 1 ) + a 2 φ ( 2 ) + ψ ,
where u i ( ρ ) , φ ( ρ ) , w 1 , and ψ are unknown functions, which are independent of x 3 , and a 1 and a 2 are unknown constants. The functions u i ( ρ ) and φ ( ρ ) are the solution of two isotermal plane strain problems, denoted by A ( ρ ) , and the functions w i and ψ are the solution of a nonisothermal plane strain problem denoted by A ( 0 ) . For brevity, the solutions of the problems A ( 0 ) and A ( ρ ) are provided directly omitting the calculations. We obtain
u 1 ( 1 ) = c 1 x 1 , u 2 ( 1 ) = c 1 x 2 , u 3 ( 1 ) = 0 , φ ( 1 ) = c 2 ,
  u 1 ( 2 ) = h 1 x 1 , u 2 ( 2 ) = h 1 x 2 , u 3 ( 2 ) = 0 , φ ( 2 ) = h 2 ,
w 1 = s 1 x 1 , w 2 = s 1 x 2 , w 3 = 0 , ψ = s 2 ,  
where
c 1 = λ ξ d 2 D , c 2 = 2 μ d D , h 1 = 2 f ξ / D , h 2 = 4 d f / D s 1 = β d b ξ D T * , s 2 = 2 [ d b β ( λ + μ ) ] D T * D = 2 [ d 2 ξ ( λ + μ ) ] .
The strain and stress tensors associated with ( u i ( 1 ) , φ ( 1 ) ) are provided by
e 11 ( 1 ) = e 22 ( 1 ) = c 1 , e 12 ( 1 ) = e 13 ( 1 ) = e 23 ( 1 ) = e 33 ( 1 ) = 0 , κ i j k ( 1 ) = 0 , τ 11 ( 1 ) = τ 22 ( 1 ) = 2 ( μ + λ ) c 1 + d c 2 , τ 12 ( 1 ) = τ 13 ( 1 ) = τ 23 ( 1 ) = 0 , μ α β γ ( 1 ) = μ α β 3 ( 1 ) = 0 , σ α ( 1 ) = 0 , g ( 1 ) = 2 c 1 d + c 2
From relations (13) we obtain
e 11 ( 2 ) = e 22 ( 2 ) = h , e 12 ( 2 ) = e 13 ( 2 ) = e 23 ( 2 ) = e 33 ( 2 ) = 0 , κ i j k ( 2 ) = 0 , τ 11 ( 2 ) = τ 22 ( 2 ) = 2 ( μ + λ ) h 1 + d h 2 , τ 12 ( 2 ) = τ 13 ( 2 ) = τ 23 ( 2 ) = 0 , μ α β γ ( 2 ) = μ α β 3 ( 2 ) = 0 , σ α ( 2 ) = 0 , g ( 2 ) = 2 h 1 d + h 2 .
The relations (14) provide
e 11 ( 0 ) = e 22 ( 0 ) = s 1 , e 12 ( 0 ) = e 13 ( 0 ) = e 23 ( 0 ) = e 33 ( 0 ) = 0 , κ i j k ( 0 ) = 0 , τ 11 ( 0 ) = τ 22 ( 0 ) = 2 ( μ + λ ) s 1 + d s 2 b T * , τ 12 ( 0 ) = τ 13 ( 0 ) = τ 23 ( 0 ) = 0 , μ α β γ ( 0 ) = μ α β 3 ( 0 ) = 0 , σ α ( 0 ) = 0 , g ( 0 ) = 2 s 1 d + s 2 β T *
We have to determine the constants a 1 and a 2 . In view of external data (10), the boundary conditions (6)–(9) became
Σ 1 P α d a + Γ 1 Q α d s = 0 ,    
Σ 1 P 3 d a + Γ 1 Q 3 d s = F 3 ,
Σ 1 ( x α P 3 + R α ) d a + Γ 1 x α Q 3 d s = 0 ,
Σ 1 ε α β 3 x α P β d a + Γ 1 ε a β 3 x α Q β d s = M 3 .
The conditions (19) and (21) in view of Equations (5), (16)–(18) are identically satisfied. Conditions (20) and (22) reduce to the following system
D 11 a 1 + D 12 a 2 = ( F 3 + F 3 * ) , D 21 a 1 + D 22 a 2 = M 3
where
D 11 = 2 π a 2 μ 3 d 2 ξ ( 2 μ + 3 λ ) D , D 12 = D 21 = 4 π a 2 f 3 d 2 ξ ( 2 μ + 3 λ ) D , D 22 = μ π a 4 2 + 4 ( 2 α 4 α 5 + 2 ξ f 2 D ) π a 2 . F 3 * = 2 μ π a 2 s 1 .
System (26) uniquely determines the constants a 1 and a 2 . We have
a 1 = [ ( F 3 + F 3 * ) D 22 M 3 D 12 ] / D * a 2 = [ ( F 3 + F 3 * ) D 12 M 3 D 11 ] / D *
where
D * = D 11 D 22 D 12 2 .
The solution (11) can be rewritten in the form
u 1 = 1 D * { F 3 [ D 12 x 2 x 3 + ( c 1 D 22 h 1 D 12 ) x 1 ] + M 3 [ D 11 x 2 x 3 + ( c 1 D 12 h 1 D 11 ) x 1 ] + γ 1 T * [ D 12 x 2 x 3 + ( c 1 D 22 h 1 D 12 + D * 2 μ π a 2 ) x 1 ] } , u 2 = 1 D * { F 3 [ D 12 x 1 x 3 + ( c 1 D 22 h 1 D 12 ) x 2 ] + M 3 [ D 11 x 1 x 3 + ( c 1 D 12 h 1 D 11 ) x 2 ] + γ 1 T * [ D 12 x 1 x 3 + ( c 1 D 22 h 1 D 12 + D * 2 μ π a 2 ) x 2 ] } , u 3 = 1 D * ( F 3 D 22 + M 3 D 12 + γ 1 T * D 22 ) x 3 φ = 1 D * { F 3 ( c 2 D 22 h 2 D 12 ) + M 3 ( c 2 D 12 h 2 D 11 ) + T * [ γ 1 ( c 2 D 22 h 2 D 12 ) + γ 2 D * ]
where
γ 1 = 2 μ π a 2 β d b ξ D , γ 2 = 2 [ d b β ( λ + μ ) ] D .
From the general solution (27), we derive the solutions of some special cases.
  • chiral circular cylinder ( b = 0 , β = 0 )
It follows from (15) and (28)
c 2 = 0 , h 2 = 0 , γ 2 = 0
so that φ = 0 . From (27) and (29), we obtain
u 1 = F 3 { 2 f μ k 1 x 2 x 3 + [ λ 2 μ ( 2 μ + 3 λ ) A + 2 f 2 μ 2 k 1 ] x 1 } + M 3 ( 1 k 1 x 2 x 3 f μ k 1 x 1 ) + T * b { 2 f A ( μ + λ ) k 1 x 2 x 3 + [ 1 2 μ + 3 λ 2 f 2 A μ ( μ + λ ) k 1 ] x 1 } u 2 = F 3 { 2 f μ k 1 x 1 x 3 + [ λ 2 μ ( 2 μ + 3 λ ) A + 2 f 2 μ 2 k 1 ] x 2 } M 3 ( 1 k 1 x 1 x 3 + f μ k 1 x 2 ) + T * b { 2 f A ( μ + λ ) k 1 x 1 x 3 + [ 1 2 μ + 3 λ 2 f 2 A μ ( μ + λ ) k 1 ] x 2 } u 3 = { F 3 [ μ + λ μ ( 2 μ + 3 λ ) A + 4 f 2 μ 2 k 1 ] + M 3 2 f μ k 1 + T * b [ 1 2 μ + 3 λ + 4 f 2 A μ ( μ + λ ) k 1 ] } x 3
where A = π a 2 , I 0 = π a 4 / 2 and
k 1 = μ I 0 + 4 ( 2 α 4 α 5 ) A 12 f 2 A μ .
  • porous circular cylinder( f = 0 )
When the material is achiral, f is equal to zero. Moreover, we put the constitutive constants associated with the strain gradient equal to zero. These assumptions imply
D 12 = 0 , D 22 = μ I 0 , D * = D 11 D 22 h 1 = h 2 = 0 .
Introducing Equations (32) into (27), we have the solution of an achiral circular cylinder in the theory of materials with voids.
u 1 = F 3 c 1 D 11 x 1 + M 3 D 22 x 2 x 3 + T * γ 1 D 11 x 1 u 2 = F 3 c 1 D 11 x 2 M 3 D 22 x 1 x 3 + T * γ 1 D 11 x 2 u 3 = ( F 3 + T * γ 1 ) 1 D 11 x 3 φ = F 3 c 2 D 11 + T * ( γ 1 c 2 D 11 + γ 2 )
where we used the relation
1 D 11 = c 1 D 11 + 1 2 μ π a 2

4. Engineering Material Constants

In this section, solution (27), (30), and (33) are rewritten in terms of technical constants.
  • porous achiral circular cylinder
We assume that the internal energy density is a positive definite form. This assumption implies that
μ > 0 , ξ > 0 , a 0 > 0 , 2 μ + 3 λ > 0 , ξ ( 2 μ + 3 λ ) > 3 d 2
We introduce the notations
E * = μ ( 2 μ + 3 λ * ) μ + λ * , ν * = λ * 2 ( μ + λ * ) ,
where
λ * = λ d 2 / ξ .
Of course
λ * < λ , E * < E , 2 μ + 3 λ * > 0 .
The elastic constants E * and ν * are the counterpart of Young modulus and Poisson ratio in the theory of elastic materials with voids. The inverse relations of (36) are provided by
2 μ = E * 1 + ν * , λ * = ν * E * ( 1 + ν * ) ( 1 2 ν * ) , 2 μ + 3 λ * = E * 1 2 ν * .
With the help of (36) and (39), we obtain
c 1 = ν * , c 2 = d ξ ( 1 2 ν * ) , γ 1 = b * ( 1 2 ν * ) A , γ 2 = 2 b 0 ( 1 + ν * ) ( 1 2 ν * ) ξ E * , D 11 = A E * , D 22 = I 0 E * 2 ( 1 + ν * ) , D = ξ E * ( 1 + ν * ) ( 1 2 ν * ) ,
where
b * = b β d ξ , b 0 = β ( λ + μ ) d b = β ( μ + λ * ) b * d .
Taking into account (40) and (41), relations (33) became
u 1 = 1 E * [ F 3 ν * A x 1 + 2 M 3 ( 1 + ν * ) I 0 x 2 x 3 + T * b * ( 1 2 ν * ) x 1 ] , u 2 = 1 E * [ F 3 ν * A x 2 2 M 3 ( 1 + ν * ) I 0 x 1 x 3 + T * b * ( 1 2 ν * ) x 2 ] , u 3 = 1 E * [ F 3 A + T * b * ( 1 2 ν * ) ] x 3 , φ = 1 2 ν * E * ξ [ F 3 d A ( 3 b * d β E * 1 2 ν * ) T * ] .
From (1) and (42), the strain tensor is provided by
e 11 = e 22 = 1 E * [ F 3 ν * A + T * b * ( 1 2 ν * ) ] , e 12 = 0 , e 33 = 1 E * [ F 3 A + T * b * ( 1 2 ν * ) ] , e 13 = M 3 ( 1 + ν * ) E * I 0 x 2 , e 23 = M 3 ( 1 + ν * ) E * I 0 x 1
From the constitutive Equations (2) and (43), we obtain
t 11 = t 22 = t 12 = 0 , t 33 = F 3 A , t 13 = M 3 I 0 x 2 , t 23 = M 3 I 0 x 1 σ 1 = σ 2 = σ 3 = 0 , g = 0
It should be mentioned that if in the Equation (42), we replace E * with E, ν * with ν , b * with b, and put φ = 0 , we obtain the solution of the analogous problem in the classical theory of elasticity.
  • chiral circular cylinder
Together with (31) and (45), the relations (30) yield
u 1 = F 3 [ 2 d 1 k 1 x 2 x 3 + ( ν E A + 2 d 1 2 k 1 ) x 1 ] + M 3 ( 1 k 1 x 2 x 3 1 d 1 k 1 x 1 ) + T * b A [ 2 d 2 k 1 x 2 x 3 + ( 1 2 ν E A 2 d 1 d 2 k 1 ) x 1 ] u 2 = F 3 [ 2 d 1 k 1 x 1 x 3 + ( ν E A + 2 d 1 2 k 1 ) x 2 ] M 3 ( 1 k 1 x 1 x 3 + 1 d 1 k 1 x 2 ) + T * b A [ 2 d 2 k 1 x 1 x 3 + ( 1 2 ν E A 2 d 1 d 2 k 1 ) x 2 ] u 3 = [ F 3 ( 1 E A + 4 d 1 2 k 1 ) + M 3 2 d 1 k 1 + T * b ( 1 2 ν E + 4 A d 1 d 2 k 1 ) ] x 3 ,
where
d 1 = E 2 ( 1 + ν ) f , d 2 = E 2 ( 1 + ν ) ( 1 2 ν ) f .
The strain tensor is provided by
e 11 = e 22 = F 3 ν E A + 2 d 1 2 k 1 M 3 2 d 1 k 1 + T * b 1 2 ν E 2 A d 1 d 2 k 1 , e 33 = F 3 1 E A + 4 d 1 2 k 1 + M 3 2 d 1 k 1 + T * b 1 2 ν E + 4 A d 1 d 2 k 1 , e 12 = 0 , e 13 = F 3 d 1 + M 3 2 T * b A d 2 x 2 k 1 , e 23 = F 3 d 1 M 3 2 + T * b A d 2 x 1 k 1
The components of the stress tensor are
t 11 = t 22 = 2 f k 1 ( 2 F 3 d 1 M 3 2 b T * A d 2 ) , t 33 = F 3 A + 2 t 11 t 13 = { E 2 ( 1 + ν ) M 3 + 2 f [ F 3 + T * b A ( 1 2 ν ) ] } x 2 k 1 t 12 = 0 , t 23 = { E 2 ( 1 + ν ) M 3 2 f [ F 3 + T * b A ( 1 2 ν ) ] } x 1 k 1
  • porous chiral circular cylinder
From (24), (26), and (40), we obtain
c 1 D 22 h 1 D 12 D * = ( ν * E * A + 2 d ^ 1 2 k 1 ) , c 1 D 12 h 1 D 11 D * = 1 d ^ 1 k 1 , c 2 D 12 h 2 D 11 D * = 0 , c 2 D 22 h 2 D 12 D * = ( 1 2 ν * ) d ξ E * A , γ 1 c 1 D 22 h 1 D 12 D * + 1 2 μ A = b * 1 2 ν * E * 2 A d 1 d 2 k 1 γ 1 c 2 D 22 h 2 D 12 D * + γ 2 = 3 ( 1 2 ν * ) b * d ξ E * + β ξ
where
d ^ 1 = E * 2 ( 1 + ν * ) f , d ^ 2 = E * 2 ( 1 + ν * ) ( 1 2 ν * ) f .
Substituting (49) into (27), we obtain
u 1 = F 3 2 d ^ 1 k 1 x 2 x 3 + ( ν * E * A + 2 d ^ 1 2 k 1 ) x 1 + M 3 1 k 1 x 2 x 3 1 d ^ 1 k 1 x 1 + T * b * A 2 d ^ 2 k 1 x 2 x 3 + 1 2 ν * E * A 2 d ^ 1 d ^ 2 k 1 x 1 u 2 = F 3 2 d ^ 1 k 1 x 1 x 3 + ν * E * A + 2 d ^ 1 2 k 1 x 2 M 3 1 k 1 x 1 x 3 + 1 d ^ 1 k 1 x 2 + T * b * A 2 d ^ 2 k 1 x 1 x 3 + 1 2 ν * E * A 2 d ^ 1 d ^ 2 k 1 x 2 u 3 = F 3 1 E * A + 4 d ^ 1 2 k 1 + 2 M 3 d ^ 1 k 1 + T * b * A 1 2 ν * E * A + 4 d ^ 1 d ^ 2 k 1 x 3 , φ = 1 2 ν * E * ξ F 3 d A + T * 3 b * d + β E * 1 2 ν * .
The strain tensor is easily determined from (51)
e 11 = e 22 = F 3 ν * E * A + 2 d ^ 1 2 k 1 M 3 d ^ 1 k 1 + T * b * A 1 2 ν * E * A 2 d ^ 1 d ^ 2 k 1 , e 33 = F 3 1 E * A + 4 d ^ 1 2 k 1 + 2 M 3 d ^ 1 k 1 + T * b * A 1 2 ν * E * A + 4 d ^ 1 d ^ 2 k 1 , e 12 = 0 , e 13 = 1 k 1 F 3 d ^ 1 + M 3 2 + T * b * A d ^ 2 x 2 , e 23 = 1 k 1 F 3 d ^ 1 + M 3 2 + T * b * A d ^ 2 x 1 .
The stress tensor is provided by
t 11 = t 22 = 2 f k 1 2 d ^ 1 M 3 2 d ^ 2 , t 33 = F 3 A 4 f k 1 2 F 3 d ^ 1 M 3 + T * b * A d ^ 2 t 13 = E * 2 ( 1 ν * ) M 3 + 2 f F 3 d ^ 1 + T * b * A ( 1 2 ν ) x 2 k 1 t 23 = E * 2 ( 1 ν * ) M 3 2 f F 3 d ^ 1 + T * b * A ( 1 2 ν ) x 1 k 1 t 12 = 0 σ i = 0 , g = 0

5. Extensional and Torsional Rigidities

In [17], Mindlin and Eshel presented three forms for the potential energy-density in the case of centrosymmetric and isotropic materials, and derived the relations connecting the three forms. In the context of non-centrosymmetric strain gradient theory, Papanicolopulos [16] modified the forms II and III by adding a term that introduced chirality in the material behaviour. In this paper, we considered form I of the strain gradient theory. The energy-density function is provided by
W = 1 2 λ e i j e i j + μ e i j + α 1 κ i i k κ k j j + α 2 κ i j j κ i k k + α 3 κ i i k κ j j k + α 4 κ i j k κ i j k + α 5 κ i j k κ k j i + 2 f ϵ i j k κ i j k e j l .
Necessary and sufficient conditions for positive definiteness of W are
μ > 0 , 2 μ + 3 λ > 0 , 2 α 4 α 5 > 0 , α 4 + α 5 > 0 , 5 3 ( α 1 + α 2 + α 3 ) + α 4 + α 5 > 0 2 ( 2 α 1 + α 2 + 4 α 3 ) + 3 ( 2 α 4 α 5 ) > 0 , 5 6 ( α 1 + 2 α 2 4 α 3 ) 2 < [ 2 ( 2 α 1 + α 2 + 4 α 3 ) + 3 ( 2 α 4 α 5 ) ] [ 5 3 ( α 1 + α 2 + α 3 ) + α 4 + α 5 ] 12 f 2 < μ ( 2 α 4 α 5 ) .
The quantity k 1 in Equation (31) is the torsional rigidity of the cylinder. It consists of three terms; the first term is the usual torsional rigidity in classical elasticity, the second term depends by the strain gradient theory, and the third negative term is due to the chirality of the material. From the last disequality of (55), it immediately follows that k 1 > 0 . In contrast with the classical theory of elasticity, each mechanical or thermal action produces extension, torsion, and a deformation in the plane of the cross section of the cylinder. As consequence, in the case under discussion, we have three torsional rigidities and three extensional rigidities. Table 1 and Table 2 show the rigidities corresponding to the forces F 3 and M 3 , and to the temperature T * .
The axial force F 3 rotates the cross-section by an angle
θ ( x 3 ) = 4 ( 1 + ν ) f E k 1 F 3 x 3 ,
in the case of a chiral cylinder, and by an angle
θ ( x 3 ) = 4 ( 1 + ν * ) f E * k 1 F 3 x 3 ,
in the case of a porous chiral cylinder. The sign of θ depends on the sign of the product F 3 f . If F 3 f > 0 , the angle θ is positive (counterclockwise), otherwise if F 3 f < 0 , θ is negative (clockwise), Figure 1a,b.
As a dual effect, the cylinder lengthens or shortens when twisted. If M 3 f > 0 , the cylinder undergoes an elongation and a contraction of the cross-section. Vice versa, if M 3 f < 0 , the torsion induces a shortening of the cylinder and a dilatation of the cross section, Figure 2a,b.
The effects of chirality vanish when f = 0 . In the case of a combined action of the axial force F 3 and the torque M 3 , the angle of torsion θ is provided by
θ ( x 3 ) = 2 F 3 d 1 M 3 x 3 k 1 .
Immediately, we see that if M 3 is related to F 3 , by the following relation
M 3 = 4 ( 1 + ν ) E f F 3 M 3 = 4 ( 1 + ν * ) E * f F 3 ,
the angle θ is equal to 0 and the torsional effect produced by F 3 vanishes. Moreover, we denote by Δ l the variation in length of the axis of the cylinder. If relation (59) holds, from (45), we obtain
Δ l = F 3 l E A Δ l = F 3 l E * A ,
and the extensional effect produced by M 3 disappears.

6. Conclusions

The results presented in this paper can be summarized as follows:
  • We present the basic equations of the strain gradient theory of chiral porous thermoelastic solids and formulate the equilibrium problem of a homogeneous and isotropic circular cylinder subjected to a prescribed axial force and a torque acting on its bases. The cylinder is also under tha action of a constant temperature field.
  • The analytical solution is determined through the help of two-dimensional problems. The solutions of a chiral (non porous) cylinder and a porous (non chiral) cylinder are derived as special cases.
  • With the introduction of suitable notations, we define engineering constants, such as Young-type modulus and Poisson-type ratio, for chiral porous materials. Explicit formulas for the displacements, microdilatation function, stresses, and strain are written in terms of such engineering constants.
  • The chirality is introduced in the constitutive equations by a material constant f. The sign of f may be positive or negative. We show that the cylinder is twisted by the axial force F 3 and the rotation will be counterclockwise if F 3 f > 0 or clockwise if F 3 f < 0 . In addition, the torque M 3 produces extension. The cylinder lengthens or shortens depending on M 3 f > 0 or M 3 f < 0 , respectively. Furthermore the cylinder is twisted by the variation in temperature T * and the sign of the product f T * discriminates between the two directions of the rotation.
  • In contrast with the classical theory of elasticity, the rigidity in the torsion of a chiral beam is measured by three moduli. Similarly, we use three moduli for the extensional rigidity. These moduli are computed and are presented in Table 1 and Table 2.
  • On the basis of the results presented in the paper, a possible next step will be to investigate the elastic deformation of a chiral porous circular beam under the action of a bending moment and a shear force. Bending and shear stiffness can be derived from the solutions of these problems.
  • We show that the bending does not occur in a chiral circular cylinder under uniaxial force, although it is predicted by the solution for cylinders with an arbitrary cross-section. It might be interesting to solve the problem of a cylinder with a non circular section to study the coupling of deformation modes using extension-bending.
  • Other important mechanical properties of chiral materials such as the ultimate tensile strength, elongation at rupture, fatigue properties, and so on need to be determined in various research areas, depending on the theoretical and experimental approaches used.

Funding

This research received no external funding.

Data Availability Statement

All data are contained within the article.

Acknowledgments

The author espresses gratitude to the referees for their helpful suggestions.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Guz, A.N.; Rushchitskii, Y. Nanomaterials: On the mechanics of nanomaterials. Int. Appl. Mech. 2003, 39, 1272–1293. [Google Scholar] [CrossRef]
  2. Lakes, R. Elastic and viscoelastic behavior of chiral materials. Int. J. Mech. Sci. 2001, 43, 1579–1589. [Google Scholar]
  3. Lakes, R.; Benedict, R.L. Noncentrosymmetry in micropolar elasticity. Int. J. Eng. Sci. 1982, 20, 1161–1167. [Google Scholar] [CrossRef]
  4. Natroshvili, D.; Stratis, I.G. Mathematical problems of the theory of elasticity of chiral materials for Lipschitz domains. Math. Methods Appl. Sci. 2006, 29, 445–478. [Google Scholar] [CrossRef]
  5. Toupin, R.A. Theories of elasticity with couple-stress. Arch. Ration. Mech. Anal. 1964, 17, 85–112. [Google Scholar] [CrossRef]
  6. Matsushima, T.; Chambon, R.; Caillerie, D. Large strain finite element analysis of a local second gradient model: Application to localization. Int. J. Numer. Methods Eng. 2002, 54, 499–521. [Google Scholar] [CrossRef]
  7. Nika, G. Cosserat continuum modelling of chiral scale-size effects and their influence on effective constitutive laws. Forces Mech. 2022, 9, 1–9. [Google Scholar] [CrossRef]
  8. Böhmer, C.G.; Lees, Y.; Neff, P. Chirality in the plane. Mech. Phys. Solids 2020, 134, 1–13. [Google Scholar] [CrossRef]
  9. Surówka, P.; Suslov, A.; Jülicher, F.; Baurjce, D. Odd Cosserat elasticity in active materials. Phys. Rev. 2023, 108, 064609. [Google Scholar] [CrossRef]
  10. Giorgio, I.; dell’Isola, F.; Misra, A. Chirality in 2D Cosserat media related to stretch-micro-rotation coupling with links to granular micromechanics. Int. J. Solids Struct. 2020, 202, 28–38. [Google Scholar] [CrossRef]
  11. Giorgio, I.; Hild, F.; Gerami, E.; dell’Isola, F.; Misra, A. Experimental verification of 2D Cosserat chirality with stretch-micro-rotation coupling in orthotropic metamaterials with granular motif. Mech. Res. Commun. 2022, 126, 104020. [Google Scholar] [CrossRef]
  12. Barchiesi, E.; dell’Isola, F.; Bersani, A.M.; Turco, E. Equilibria determination of elastic articulated duoskelian beams in 2D via Riks-type algorithm. Int. J. Non-linear Mech. 2021, 128, 103628. [Google Scholar] [CrossRef]
  13. Misra, A.; Nejadsadeghi, N.; De Angelo, M.; Placidi, L. Chiral metamaterial predicted by granular micromechanics: Verified with 1D example synthesized using additive manifacturing. Cont. Mech. Thermodyn. 2020, 32, 1497–1513. [Google Scholar] [CrossRef]
  14. Zhang, Y.Q.; Lin, G.R.; Wang, J.S. Small-scale effects on buckling of multiwalled carbon nanotubes under axial compression. Phys. Rev. B 2004, 70, 205430. [Google Scholar] [CrossRef]
  15. Duan, W.; Wang, Q.; Liew, K.; He, X. Molecular mechanics modelling of carbon nanotube fracture. Carbon 2007, 45, 1769–1776. [Google Scholar] [CrossRef]
  16. Papanicolopulos, S.A. Chirality in isotropic linear gradient elasticity. Int. J. Solids Struct. 2011, 48, 745–752. [Google Scholar] [CrossRef]
  17. Mindlin, R.D.; Eshel, N.N. On first strain gradient theories in linear elasticity. Int. J. Solids Struct. 1968, 4, 109–112. [Google Scholar] [CrossRef]
  18. Lu, G.Q.; Zhao, X.S. Nanoporous Materials. Science and Engineering; Series on Chemical Engineering 4; Imperial College Press: London, UK, 2004. [Google Scholar]
  19. Ernst, S. (Ed.) Advances in Nanoporous Materials; Elsevier: Amsterdam, The Netherlands, 2009. [Google Scholar]
  20. Svanadze, M. Potential Method in Mathematcal Theories of Multi-Porosity Media; Springer International Publishing: Cham, Switzerland, 2019. [Google Scholar]
  21. Ieşan, D. A theory of thermoelastic materials with voids. Acta Mech. 1986, 60, 67–89. [Google Scholar] [CrossRef]
  22. De Cicco, S. Non-simple elastic materials with double porosity structure. Arch. Mech. 2022, 74, 127–142. [Google Scholar]
  23. Eringen, A.C. Microcontinuum Field Theories. Foundations and Solids; Springer: Berlin/Heidelberg, Germany; New York, NY, USA, 1998. [Google Scholar]
  24. Ieşan, D. Thermoelastic Models of Continua; Kluwer Academic Publishers: Dordrecht, The Netherlands; Boston, MA, USA; London, UK, 2004. [Google Scholar]
  25. Cowin, S.C.; Nunziato, J.W. Linear elastic materials with voids. Elasticity 1983, 13, 125–147. [Google Scholar] [CrossRef]
  26. De Cicco, S.; De Angelis, F. A plane strain problem in the theory of elastic materials with voids. Math. Mech. Solids 2020, 25, 46–59. [Google Scholar] [CrossRef]
  27. Svanadze, M.; De Cicco, S. Fundamental solutions in the full coupled linear theory of elasticity for solids with double porosity. Arch. Mech. 2013, 65, 367–390. [Google Scholar]
  28. De Cicco, S.; Ieşan, D. On the theory of thermoelastic materials with a double porosity structure. Therm. Stress. 2021, 44, 1514–1533. [Google Scholar] [CrossRef]
  29. De Cicco, S. On the deformation of porous spherical bodies under radial surface traction. J. Theor. Appl. Mech. 2023, 61, 5. [Google Scholar] [CrossRef]
  30. Sciarra, G.; Dell’Isola, F. Second gradient poromechanics. Int. J. Solids Struct. 2007, 47, 6607–6629. [Google Scholar] [CrossRef]
  31. Dell’Isola, F.; Gavrilyuk, S. Variational Models and Methods in Solid and Fluid Mechanics; CISM Courses and Lectures 535; Springer: Berlin/Heidelberg, Germany, 2012. [Google Scholar]
  32. De Cicco, S. On the deformation of porous spherical shells and other spherical bodies. Current perspectives and new directions in mechanics modelling and design of structural systems. In Proceedings of the 8th International Conference on Structural Engineering, Mechanics and Computation, Cape Town, South Africa, 5–7 September 2022; CRC Press: Boca Raton, FL, USA, 2022; pp. 309–314. [Google Scholar]
  33. Ellouz, H.; Jrad, H.; Wali, M.; Dammak, F. Large deflection analysis of FGM/magneto-electro-elastic smart shells with porosities under multi-physics loading. Mech. Adv. Mater. Struct. 2023, in press. [Google Scholar] [CrossRef]
  34. Auffray, N.; Bouchet, R.; Brechet, Y. Derivation of anisotropic matrix for bidimensional strain-gradient elasticity behaviour. Int. J. Solids Struct. 2009, 46, 440–454. [Google Scholar] [CrossRef]
  35. Ieşan, D.; Quintanilla, R. On chiral effects in strain gradient elasticity. Eur. J. Mech. A Solids 2016, 58, 233–246. [Google Scholar] [CrossRef]
  36. Ieşan, D. Classical and Generalized Models of Elastic Rods; CRC Press: Boca Raton, FL, USA, 2009. [Google Scholar]
  37. Dell’Isola, F.; Batra, R.C. Saint-Venant’s problem for porous linear elastic materials. Elasticity 1997, 47, 7–81. [Google Scholar]
  38. Ghiba, I.D. On the deformation of transversely isotropic porous elastic circular cylinder. Arch. Mech. 2009, 5, 407–421. [Google Scholar]
  39. Birsan, M.; Altenbach, H. Theory of thermoelastic rods made of porous materials. Arch. Appl. Mech. 2011, 81, 1365–1391. [Google Scholar] [CrossRef]
  40. De Cicco, S.; Ieşan, D. Thermal effects in anisotropic porous elastic rods. Therm. Stress. 2013, 81, 364–377. [Google Scholar] [CrossRef]
  41. Malikan, M.; Eremeyev, V.A.; Zur, K.K. Effect of Axial Porosities on Flexomagnetic Response of In-Plane Compressed Piezomagnetic Nanobeams. Symmetry 2020, 12, 1935. [Google Scholar] [CrossRef]
  42. Malikan, M.; Eremeyev, V.A. On the Dynamics of a Visco–Piezo–Flexoelectric Nanobeam. Symmetry 2020, 12, 643. [Google Scholar] [CrossRef]
  43. Arda, M.; Aydogdu, M. Torsional dynamics of coaxial nanotubes with different lengths in viscoelastic medium. Microsyst. Technol. 2019, 25, 3943–3957. [Google Scholar] [CrossRef]
  44. Akgözn, B.; Civalek, O. Bending analysis of embedded carbon nanotubes resting on an elastic foundation using strain gradient theory. Acta Astronaut. 2016, 119, 1–12. [Google Scholar] [CrossRef]
  45. Akgözn, B.; Civalek, O. A novel microstructure-dependent shear deformable beam model. Int. J. Mech. Sci. 2015, 99, 10–20. [Google Scholar] [CrossRef]
  46. Billington, E.W. The poynting effect. Acta Mech. 1986, 58, 19–31. [Google Scholar] [CrossRef]
  47. Dell’Isola, F.; Ruta, G.C.; Batra, R.C. Generalized poynting effects in predeformed prismatic bars. Elasticity 1998, 50, 181–196. [Google Scholar] [CrossRef]
  48. Horgan, C.O.; Murphy, J.G. Poynting and reverse poynting effects in soft materials. Soft Matter 2017, 28, 4916–4923. [Google Scholar] [CrossRef]
  49. De Cicco, S.; Ieşan, D. On the thermal stresses in chiral porous elastic beams. Contin. Mech. Thermodyn. 2023, 35, 2095–2115. [Google Scholar] [CrossRef]
Figure 1. (a) Torsion of a chiral circular beam under compression; (b) Torsion of a chiral circular beam under a traction force.
Figure 1. (a) Torsion of a chiral circular beam under compression; (b) Torsion of a chiral circular beam under a traction force.
Symmetry 16 00129 g001
Figure 2. (a) Extension of a chiral circular beam under clockwise torque (on Σ 2 ; (b) extension of a chiral circular beam under counterclockwise torque (on Σ 2 ).
Figure 2. (a) Extension of a chiral circular beam under clockwise torque (on Σ 2 ; (b) extension of a chiral circular beam under counterclockwise torque (on Σ 2 ).
Symmetry 16 00129 g002
Table 1. Rigidities of the chiral circular cylinder.
Table 1. Rigidities of the chiral circular cylinder.
Torsional RigidityExtensional Rigidity
F 3 1 2 d 1 k 1 1 E A + 4 d 1 2 k 1 1
M 3 k 1 1 2 d 1 k 1
T * d 2 k 1 2 b A 1 b 1 2 ν E + 4 A d 1 d 2 k 1 1
Table 2. Rigidities of chiral porous circular cylinder.
Table 2. Rigidities of chiral porous circular cylinder.
Torsional RigidityExtensional Rigidity
F 3 1 2 d ^ 1 k 1 1 E * A + 4 d ^ 1 2 k 1 1
M 3 k 1 1 2 d ^ 1 k 1
T * d ^ 2 k 1 2 b A 1 b 1 2 ν * E * + 4 A d ^ 1 d ^ 2 k 1 1
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

De Cicco, S. Explicit Formulas for the Deformation of Chiral Porous Circular Beams in Gradient Thermoelasticity. Symmetry 2024, 16, 129. https://doi.org/10.3390/sym16010129

AMA Style

De Cicco S. Explicit Formulas for the Deformation of Chiral Porous Circular Beams in Gradient Thermoelasticity. Symmetry. 2024; 16(1):129. https://doi.org/10.3390/sym16010129

Chicago/Turabian Style

De Cicco, Simona. 2024. "Explicit Formulas for the Deformation of Chiral Porous Circular Beams in Gradient Thermoelasticity" Symmetry 16, no. 1: 129. https://doi.org/10.3390/sym16010129

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop