Skip to main content

Open Access 25.04.2024 | Original

A dislocation-based finite element method for plastic collapse assessment in solid mechanics

verfasst von: Renato Zona, Vincenzo Minutolo

Erschienen in: Archive of Applied Mechanics

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

A displacement-based dislocation map has been used to build the eigenstress stress, which is the base of the structure’s limit analysis. The limit load has been calculated as the upper bound of any equilibrated stress that respects the compatibility inequalities by means of a linear optimization program. The eigenstress stress nodal parameters were assumed as the design variables, and the compatibility inequalities have been obtained from the Mises–Schleicher criterion, assuming that the stress belongs to the corresponding plastic domain. The numerical application has considered a linear secant representation of the domain, with a penalty factor on stresses, to correct the linearization error. Examples concerning a simply supported cantilever beam, a pipe section, and a plate with a circular hole highlighted the accuracy of the procedure with respect to the established literature. Moreover, the procedure has been applied to investigate plane structure examples. A square plate with variable elliptic holes has been analyzed, and the influence of ellipticity on the collapse load has been shown. The effects of porosity and heterogeneity of the structure with respect to the collapse load are shown considering the porous cantilever and representative volume element. The evaluation of the limit load along different element directions envisaged a point-wise calculation of the compatibility domain of the porous material to be used in the macro-scale analysis of the structures made of porous micro-cells.
Hinweise

Publisher's Note

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

1 Introduction

In most structural analyses, the limit load and the required ultimate displacement before the loss of functionality are the desired results from the structural calculation. Accurate time-history analysis of structures is possible only when the load path to which the structure is subjected is known in detail. In the majority of practical applications, on the contrary, the load path is unknown, and only the load intensity range is fixed by statistical evaluation of the bounds of the load intensity, see, for example, the American Society of Civil Engineers documentation [1]. The limit analysis allows calculating the ultimate load factor of the structures from the knowledge of the load pattern and the structural geometric and constitutive properties, even if the load time history is not known. Indeed, the load safety threshold calculation is based on the knowledge of the domain of the loads, whose dimensions are defined by the load multiplier, i.e., the locus to which the loads must belong to satisfy the material strength requirements. To apply the limit design to the structural calculation through the finite element method (FEM), it is crucial to define the way one can represent the set of the eigenstress in the structures as a linear span of discrete parameters. The eigenstress set is the null space of the equilibrium operator of the structure and is related to eigenstrain, which represents any imposed deformation depending on anelasticity [2].
The application of limit analysis and plasticity to structural safety concerns a wide range of engineering fields [3]. It was the main topic of Prager, Drucker, and Greenberg’s works [48]. Massonnet and Save [9] address the plastic response of structures introducing the collapse calculation for one-dimensional beam assembly and constitute the starting point of several researches [10].
A twofold approach to the limit analysis is found in the literature. The first, the kinematic method, consists in finding the collapse load as the lower bound among the loads in equilibrium with the stress according to a compatible collapse mechanism. The second, the static approach consists in finding the upper bound among the loads in equilibrium with one of the admissible stresses (defined as the combination of one of the eigenstress stresses and the stress in equilibrium with the loads under the elastic hypothesis).
In [11, 12], the plasticity of structures has been widely studied in primis with reference to monotonic load paths. In [1316], the load path is assumed to be randomly variable, so the shakedown is addressed using mathematical programming algorithms. In soil mechanics, the limit analysis through the plasticity methods is widely applied; see Chen [17]. In the works [18, 19], the lower-bound and upper-bound theorems for geo-mechanics are formulated.
In the fields of porous media and bio-mechanics, the limit equilibrium is used as a primary tool to assess the fracture risk of prosthesis implants and corresponding optimization strategies as reported in [2023].
The numerical formulation of the plasticity method can be found in the books [2426], where iterative step-by-step analyses through the finite elements are discussed in detail. Strain concentration and large displacement in pipe limit analysis addressed via variational approach can be found in [27]. The book [28] is entirely devoted to the inelastic behavior of structures, where the algorithms of computational plasticity are described. In [2932], an iterative method is proposed to calculate the shakedown of elastic-perfectly-plastic structures and a fast incremental-iterative solution method suitable for the FEM analyses of large structures. Furthermore, [33], using the boundary element method, evaluates the eigenstress stress through an iterative algorithm and integral equations.
A semi-analytical approach is proposed in [34, 35] where the eigenstress stress for axial symmetrical domes is evaluated, and a comparison with the experimental campaign and crack lines analytical calculations reported in [36, 37] is discussed highlighting the good confidence of the method with experiments. There are several procedures for calculating the limit load of structures in terms of both one-dimensional and two-dimensional limit analysis based on the assumption that the material constituting the structures has no resistance to tension. The applicability of the limit analysis to this family of materials, called no-tension materials, i.e., concrete, stone, soils, and ceramics, can be found in [38], where a complete theoretical framework is provided. Many pioneering works deal with the limit analysis of no-tension materials [39, 40], which has been subsequently implemented through the finite elements, [4144] and [45]. In [46, 47], a critical evaluation of state-of-the-art computational methods concerning variational formulations, selection of interpolation spaces, numerical solution strategies, and numerical stability. Although these arguments are interconnected, a topic-oriented presentation is employed to thoroughly examine the abundant available literature. Lastly, the performance of several significant mixed finite element formulations is investigated in numerical simulations. The automated computation of limit analysis problem is described in [48] The optimization problem is solved numerically using the interior-point solver Mosek, taking advantage of the conic representation for yield criteria. The paper demonstrates the effectiveness of the approach through illustrative examples, spanning classical continuum limit analysis problems to generalized mechanical models. . A class of strain–displacement elements in upper-bound limit analysis is presented in [49]. The paper stated that only linear and quadratic displacement elements with straight sides, generating constant and linear strains, meet specific requirements. The text theoretically and practically demonstrates the existence of a broader class of triangular elements with straight sides, based on Bernstein polynomials, capable of providing rigorous upper bounds. This class allows for polynomial orders without upper restrictions and accommodates both continuous and discontinuous displacements. In [50], a non-incremental energy-based algorithm is proposed to calculate the equivalent material’s distribution and orientation, to achieve a compression-only state of stress. A finite element procedure where the structure is discretized using rigid infinitely resistant eight-node hexahedral elements and nonlinear interfaces, exhibiting deterioration of the mechanical properties, is proposed in [51].
The aim of the present work is to propose a general procedure for formulating the limit design through the FEM. The proposed formulation assumes that the strain parameters govern the phenomenon. The method is based on the calculation of the limit load for collapse or shakedown through the permanent strain. To this scope, Volterra’s dislocation, from which the name of the procedure, VFEM, at the element level has been assumed to represent the permanent strain, which has been described numerically by displacement-like coordinates. The whole procedure results in the definition of a numerical linear operator named the dislocation matrix, which maps the nodal dislocation parameters and the constraint displacements on the eigenstress stress. The proposed representation does not depend on the constitutive equation of the material. Furthermore, it can be seen that in the analyzed cases the principal values of the dislocation matrix are related to the multiplicity of the equilibrium equation solutions. The proposed representation is useful to generalize the construction of the eigenstress stress space in the FEM and the implementation of the lower bound theorem of the limit analysis, both for monotonic and random variable loads within prescribed limits [26, 52]. In the following, the finite element formulation of the dislocation-based limit analysis is described the eigenstress stress representation in terms of dislocations and constraint is derived, and the constrained optimization problem is described. The procedure has been evaluated through comparison with standard examples from the literature. The proposed procedure implements the direct application of the lower bound theorem of the limit analysis in the Mélan form. It allows the calculation of the collapse and the shakedown limits of a structure under several load patterns through pure elastic calculation. The key aspect of the proposed method is to generalize the static theorem application as used for framed structures and trusses to any discretization through finite elements of articulated 2D and 3D structures. Moreover, the procedure is based on elastic calculations only and does not require step by step, iterative, or predictor-corrector strategies.
In the following, 2D structures have been considered. Namely, a Holed Plate with tensile loads, a cantilever beam with different geometry, a clamped simply-supported beam with a transversal load, and a pipe annular cross-section subjected to internal pressure. The method has also been applied to porous media modeled as a clamped plate with several random holes. Finally, the numerical evaluation of the strength of a porous elementary volume has been addressed.

2 FEM for the limit analysis of structures

This section presents a procedure to evaluate the collapse and the shakedown load of structures using FEM representation of eigenstress, i.e., the residual stress due to dislocation arising. The procedure uses a finite element description of Volterra’s dislocation. The approach is devoted to linking the permanent strain that arises from the ultra-elastic behavior of the structure to the nodal parameters governing the growth of dislocations. In particular, it is shown hereafter that the permanent strain and the corresponding eigenstress stress can be represented in terms of pseudo-displacement and constraint nodal parameters. Calculating the eigenstress stress allows the formulation of the limit load assessment algorithm in terms of eigenstress according to Melan’s theorem.

2.1 FEM representation of eigenstress stress

The structure is divided into ne elements, as usual. The vector \(\textbf{u}(\textbf{x})\) represents the displacement field of the structure. Shape functions \({\mathbf {\phi }}\left( \textbf{x}\right) \) map the displacements through nodal values \(\textbf{v}_e\) on nodes, which are the parameters of the deformation. The subscript e indicates the element to which the parameters belong. The representation of the displacement is:
$$\begin{aligned} {\textbf{u}}\left( \textbf{x}\right) ={\mathbf {\phi }}\left( \textbf{x}\right) \textbf{v}_e \end{aligned}$$
(1)
The deformation is represented by the vector \(\varvec{\epsilon }_e\):
$$\begin{aligned} \varvec{\epsilon }_e=\textbf{B}_e\textbf{v}_e \end{aligned}$$
(2)
where
$$\begin{aligned} \textbf{B}_e=\mathcal {D}_\partial {\varvec{\phi }\left( \textbf{x}\right) } \end{aligned}$$
(3)
depends on the differential linear representation of the strain.
The key assumption of the proposal is that permanent strain is assumed to have the same representation as elastic deformation. Hence, the permanent strain interpolation depends on permanent nodal parameters \(\textbf{v}_e^p\) that play the role of a non-conforming nodal displacement:
$$\begin{aligned} \varvec{\epsilon }_e^p=\textbf{B}_e\textbf{v}_e^p \end{aligned}$$
(4)
consequently, the linear elastic stress–strain relationship becomes:
$$\begin{aligned} \varvec{\sigma }_e={\mathbb {C}}({\textbf{B}_e}\textbf{v}_e- {\textbf{B}_e}\textbf{v}_e^p) \end{aligned}$$
(5)
where \(\mathbb {C}\) stand for the elastic constants matrix. Introducing the stress matrix \(\textbf{K}_{e\sigma }=\mathbb {C}\textbf{B}_e\) the stress vector \(\varvec{\sigma }_e\) to be represented in terms of nodal parameters through:
$$\begin{aligned} \varvec{\sigma }_e=\textbf{K}_{e\sigma }(\textbf{v}_e-\textbf{v}_e^p) \end{aligned}$$
(6)
.
A typical example of the element dislocation map with the proposed parametrization can be seen in Fig. 1 regarding a quadrilateral four-node Lagrangian plane element.

2.2 Balance equations

The balance equation for the whole discrete structures has been obtained by stationarity of total potential energy with respect to configuration variation \(\delta \textbf{v}_e\):
$$\begin{aligned} \begin{gathered} \sum _{ne}{\delta \textbf{v}_e^T\left[ \int _{\Omega _e}{\left( \textbf{B}_e^T\mathbb {C}\left( \textbf{B}_e\textbf{v}_e-\textbf{B}_e\textbf{v}_e^p\right) -\varvec{\phi }\textbf{b}\right) \textrm{d}V-\varvec{Q}_e}\right] }=0,\ \forall \delta \textbf{v}_e^T \implies \\ \sum _{ne}\left[ {\int _{\Omega _e}{\left( \textbf{B}_e^T\mathbb {C}\left( \textbf{B}_e\textbf{v}_e-\textbf{B}_e\textbf{v}_e^p\right) -\varvec{\phi }\textbf{b}\right) \textrm{d}V-\varvec{Q}_e}}\right] =0 \end{gathered} \end{aligned}$$
(7)
where the apex T indicates transposition, \(\textbf{Q}_e\) is the element equivalent nodal force, \(\textbf{b}\) is the body force acting on the element, \(\Omega _e\) is the element domain. Equation (7) modifies with the introduction of the stiffness matrix \(\textbf{K}_e\) of the element obtained by the integration of the deformation matrix \(\textbf{B}_e\) together with the constitutive elastic matrix \(\mathbb {C}\) [24]:
$$\begin{aligned} \textbf{K}_e=\int _{\Omega _e}{\textbf{B}_e^T\mathbb {C}\textbf{B}_e\textrm{d}V} \end{aligned}$$
(8)
and of the unbalanced nodal forces on the element given by
$$\begin{aligned} \textbf{F}_e=\int _{\Omega _e}{\varvec{\phi }\textbf{b}\,\textrm{d}V+\textbf{Q}_e} \end{aligned}$$
(9)
The nodal forces and the displacement parameters of the elements assemble into the vectors of unbalanced nodal forces \(\textbf{F}=\bigcup \limits _{e} \textbf{F}_e\), nodal displacements \(\textbf{v}=\bigcup \limits _{e} \textbf{v}_e\), and nodal dislocation parameters \(\textbf{v}^p=\bigcup \limits _{e} \textbf{v}^p_e\). Furthermore, the stiffness matrices are collected in a global stiffness matrix \(\textbf{K}\). The balance equation has the following numerical form involving the applied forces and the nodal values of displacement in the unbalanced and permanent form:
$$\begin{aligned} \textbf{F}=\textbf{Kv}-\textbf{K}\textbf{v}^p \end{aligned}$$
(10)
The compatibility of the displacement within the structure is governed by the matrix \({\varvec{\mathcal {A}}}\) that expresses the element-wise displacements \(\textbf{v}\) at the nodes as a function of the local representation \(\textbf{a}\):
$$\begin{aligned} \textbf{v}={\varvec{\mathcal {A}}}\textbf{a} \end{aligned}$$
(11)
The equilibrium operator that ensures the nodal force balance is the transpose matrix \(\varvec{\mathcal {A}}^T\); consequently, the balanced nodal forces are:
$$\begin{aligned} {\varvec{\mathcal {A}}}^T\textbf{F}={\varvec{\mathcal {A}}}^T\textbf{K}{\varvec{\mathcal {A}}}\textbf{a}-{\varvec{\mathcal {A}}}^T\textbf{K}\textbf{v}^p \end{aligned}$$
(12)
In (12), the global assembled matrices can be introduced: the assembled stiffness \(\textbf{S}={\varvec{\mathcal {A}}}^T\textbf{K}{\varvec{\mathcal {A}}}\), the initial deformation matrix \(\textbf{H}={\varvec{\mathcal {A}}}^T\textbf{K}\), and the global balanced nodal forces, \(\varvec{\mathcal {F}}={\varvec{\mathcal {A}}}^T\textbf{F}\). The equilibrium equation at nodes results:
$$\begin{aligned} \varvec{\mathcal {F}}=\textbf{Sa}-\textbf{H}\textbf{v}^p \end{aligned}$$
(13)
Equation (13) contains the equilibrium and kinematic compatibility of the structure.

2.3 Eigenstress stress base

Equation (13) is the equilibrium equation that relates all the applied forces to the nodal displacements and dislocation parameters. Well-posedness of the boundary value problem requires that the parts of the vector of the prescribed displacement \(\textbf{a}_c \in \textbf{a}\) and applied force \(\varvec{\mathcal {F}}_a \in \varvec{\mathcal {F}}\) are orthogonal sets, and so are the free parts \(\textbf{a}_f=\textbf{a}\backslash \textbf{a}_c\) and \(\varvec{\mathcal {F}}_r=\varvec{\mathcal {F}}\backslash \varvec{\mathcal {F}_a}\).
The eigenstress solution of the structure corresponds to vanishing applied forces, \(\varvec{\mathcal {F}}_a=\textbf{0}\). Thus, it is possible to obtain, by exploiting Eq. (13), the unknown displacements \(\textbf{a}_f\) in terms of the displacement constraints \({\varvec{a}}_c\) and the dislocation parameters \(\textbf{v}^p\). Consequently, substituting (11) in (6) the eigenstress stress \(\varvec{\sigma }^0\) can be expressed as a linear function of the parameter \(\varvec{\delta }=\textbf{a}_c \cup \textbf{v}^p\). The whole procedure provides the dislocation matrix \(\varvec{\mathcal {V}}\) governing the relationship
$$\begin{aligned} \varvec{\sigma }^0=\varvec{\mathcal {V}}\varvec{\delta } \end{aligned}$$
(14)
The detailed closed-form expression of \(\varvec{\mathcal {V}}\) is omitted for brevity. However, a procedural technique allows straightforward calculation of it.
In all the numerical simulations, the rank of \(\varvec{\mathcal {V}}\) is equal to the structure’s redundancies that coincide with the solution’s multiplicity of the balance equations. For trusses and frames, it is finite and does not depend on discretization; for other structures, it depends on discretization. However, the rank of \(\varvec{\mathcal {V}}\) is generally less than its dimensions. The points where the stress is calculated can differ from nodes, depending on the choice of the sampling points where stress compatibility is enforced. As a consequence, the number of \(\varvec{\mathcal {V}}\) rows can vary arbitrarily. Conversely, the number of columns of \(\varvec{\mathcal {V}}\) depends on the discretization, and it is linked to the discretized structure’s mechanics and redundancy. As a consequence of the ratio between dimensions and rank of the matrix \(\varvec{\mathcal {V}}\), Eq. (14) is not invertible. The representation (14), however, spans the space of eigenstress stress in the discretized structure as a function of \(\varvec{\delta }\), so that any equilibrated stress distribution \(\varvec{\sigma }_l\) can be represented as the superimposition of the map (14) with the elastic solution to the actual loads \(\varvec{\sigma }_e\):
$$\begin{aligned} \varvec{\sigma }_l=\varvec{\sigma }_e+\varvec{\mathcal {V}} \varvec{\delta } \end{aligned}$$
(15)

2.4 Limit analysis as mathematical programming

The procedure assumes that if a constant vector \(\varvec{\bar{\delta }}\) exists, which ensures that the stress belongs to the elastic domain at any step of the load path, the structure does not undergo an unbounded increment of plastic strain. Hence, it does not collapse. The load intensity has been assumed to depend on a scalar parameter k that increases in time and multiplies one or several load patterns. In the case of one single load pattern, the collapse phenomenon occurs at a limit value \(s_c\) of k, which is called the collapse multiplier. If several load patterns vary randomly, when the plastic strain accumulates indefinitely, either ratcheting or alternate plasticity occurs, and the safety with respect to such evidence corresponds to the shakedown limit \(s_{sd}\) of k. It has to be stressed that in the case of multiple load conditions leading to shakedown, any multiplier lower than \(s_{sd}\) and greater than the minimum of the elastic multipliers is a shakedown multiplier. The mathematical program to calculate the limit multiplier to collapse or to shakedown is formulated as follows:
$$\begin{aligned} s_\alpha ={\sup _{\varvec{\delta }}} \left. k\right| f\left( k\varvec{\sigma }_e +\varvec{\mathcal {V}}\varvec{\delta }\right) \le 0 \end{aligned}$$
(16)
where \(\alpha \) stands for sd or c depending on the considered case of shakedown or collapse.
The function f in Eq. (16) represents the domain in the stress space that expresses the compatibility constraints and depends on the material constitutive law and characterization. Different constitutive laws give rise to different constraints of the optimization program.

2.5 Linearization of the constitutive equation

The proposed formulation has been applied to two-dimensional structures in a plane stress state. The analysis involves a plate in the xy plane. The stress in the plate is described by the Cauchy plane stress tensor:
$$\begin{aligned} \left[ \begin{matrix} \sigma _x &{}\quad \tau _{xy} &{}\quad 0\\ \tau _{xy} &{}\quad \sigma _y &{}\quad 0\\ 0 &{}\quad 0 &{}\quad 0 \end{matrix} \right] \end{aligned}$$
(17)
The Mises–Schleicher criterion for plane stress [26] holds for the material of the plate. It can take into account the possibility of different tension and compression limit stress, \(\sigma ^+\) and \(\sigma ^-\), respectively. In the stress space, the stress compatibility is governed by inequality:
$$\begin{aligned} \sigma _x^2 + \sigma _y^2 - \sigma _x \sigma _y + 3 \tau _{xy}^2 - (\sigma ^+ + \sigma ^-) (\sigma _x + \sigma _y) + \sigma ^+ \sigma ^- \le 0 \end{aligned}$$
(18)
that reduces to the Mises criterion, provided that the tension and compression limit stress are opposite. The inequality (18) defines a domain that constitutes the constraint function in optimization program (16). The compatibility inequalities arising from the specification of the domain to actual cases are nonlinear. In several works, the nonlinear optimization programming is addressed in the limit analysis framework [53, 54]. Iterative strategies using the quasi-Newtonian method are proposed. In particular, point-wise tangent representation of the compatibility domain is used to enforce step-by-step compatibility.
In the following calculations, a linear programming algorithm has been used to apply the limit analysis directly, avoiding iterative calculations. The domain has been linearized through the octahedron inscribed in the Schleicher ellipsoidal domain, formed by the secant planes, and that coincides with the nonlinear locus at the axis intercepts. The stress vector \(\left\{ \sigma _x, \sigma _y, \tau _{xy} \right\} \) satisfies the compatibility condition when it belongs to the octahedron interior:
$$\begin{aligned} \sigma _x \sigma _y^0 \tau _{xy}^0 + \sigma _x^0 \sigma _y \tau _{xy}^0+\sigma _x^0 \sigma _y^0 \tau _{xy} \le \sigma _x^0 \sigma _y^0 \tau _{xy}^0 \end{aligned}$$
(19)
The relationship (19) is a set of eight inequalities depending on the stress intercepts \(\sigma _x^0\), \(\sigma _y^0\), and \(\tau _{xy}^0\) with the reference axis in the \(\left\{ \sigma _x, \sigma _y, \tau _{xy} \right\} \) stress space, where the limits assume the different values, depending on the considered octahedral plane:
$$\begin{aligned} \begin{matrix} \sigma _x^0&{} = {\left\{ \begin{array}{ll} \sigma ^+ \\ \sigma ^- \end{array}\right. } &{} \sigma _y^0&{}= {\left\{ \begin{array}{ll} \sigma ^+\\ \sigma ^- \end{array}\right. } &{} \tau _{xy}^0&{}= {\left\{ \begin{array}{ll} +\sqrt{\frac{-\sigma ^+\sigma ^-}{3}}\\ -\sqrt{\frac{-\sigma ^+\sigma ^-}{3}} \end{array}\right. } \end{matrix} \end{aligned}$$
(20)
The domain and its linearization are sketched in Fig. 2. The linearization causes the constraint inequalities to underestimate the material strength except on the frame axes, namely for uniaxial stress, and the obtained limit load is less than the true collapse load. A corrector of the procedure is proposed by introducing a penalty factor to the stress intensity at any point, which corresponds to the ratio between the distance of the stress vector from the ellipsoid domain boundary and the tetrahedron one. A schematic 2D representation of the penalty factor is represented in Fig. 3 in the \(\sigma _x\),\(\tau _{xy}\) plane.

3 Numerical example

In the present section, some examples have been analyzed considering the 2D model of plates. The structures have been discretized using Lagrange’s elements with four nodes and linear interpolation of the displacement. The proposed examples presented no evidence of convergence problems due to the absence of slenderness effects, nonlinear geometric effects, and constitutive nonlinearity. The constraints of the optimization program have been written in terms of the corresponding stress components. Equation (16) contains only elastic quantities calculated at nominal load intensity. Neither elastic–plastic results are required nor iterative predictor–corrector strategies. The whole process reduces to a constrained maximum problem that is solved using mathematical programming algorithms. In the examples, ©MATLAB optimization tools have been used, and in particular linear programming has been implemented through the linearization of the inequalities described above. The linearization allowed us to simplify the solution and apply simple convergent algorithms.

3.1 Clamped-simply supported Beam

The first limit load calculation concerned the comparison of the proposed formulation with standard step-by-step commercial FEM analysis. The structure here considered is a clamped-simply supported beam, as shown in Fig. 4, with the ratio between its length and height equal to 0.1. Mechanical constraints are one fixed end and another simply supported. The beam is loaded by uniform vertical force per unit length P. The material has been assumed to be iso-resistant at the limit, hence \(\sigma ^+=-\sigma ^-=\sigma _0\). The structure has been calculated using \(\mathcal {V}FEM\), where the FEM structure discretization used four-node quadrilateral elements and Lagrange shape functions. The material constituting the structure was linearly elastic and perfectly plastic; the elastic modulus, \(E=2.1\cdot 10^5\)MPa, and Poisson ratio, \(\nu =0.3\) do not influence the results. The limit domain is described in the stress space by Eq. (19) with the position (20) that assume a linearized limit domain, see Fig. 2. The calculation of the collapse load \(P_\textrm{lim}\) has been performed either by applying the penalty factor to account for the nonlinearity of the domain or by not applying it. Table 1 reports the collapse multiplier \(s_c=P_\textrm{lim}/\sigma _0\) calculated with several element mesh numbers through the proposed formulation with and without the penalty. A step-by-step calculation of the collapse load using a FEM commercial code, and a mesh of 640 elements, has been considered for comparison, which represents a reference solution whose resulting collapse multiplier is:
$$\begin{aligned} s_{c}^a=\frac{P_{lim}^a}{\sigma _0}=3.11 \cdot 10^{-2} \end{aligned}$$
(21)
.
Two main considerations are naturally emerging from the observation of the results. First, the use of the penalty factor in the constraint inequalities produces a greater collapse multiplier, as was expected, since the nonlinearity of the solution’s domain. Second, the results have shown that the collapse multiplier obtained using the linear domain without penalty correction depends on the mesh size. This result seems to depend on the fact that the distance of the actual stress from the linearized limit is constant. Conversely, the distance from the nonlinear limit domain varies depending on the direction of the actual stress. Hence, the mesh refinements see 5 that produce the calculation of the stress at new nodes produce a slight rotation of the stress vector at the element level that has influenced the constraints of the optimization program.
Assuming the ratio between the difference of reference and actual load over reference load as an error measure of the calculation, one can verify that for the finest mesh:
$$\begin{aligned} \frac{|s_{c}^a-s_{s}|}{s_{s}^a}=\left\{ \begin{array}{ll} \text {No penalty} &{}0.14 \\ \text {Penalty} &{}0.042 \\ \end{array} \right. \end{aligned}$$
(22)
Table 1
Clamped-simply-supported beam collapse multiplier vs. mesh refinement
  
\(\varvec{\mathcal {V}}FEM\) \(s_c \cdot 10^{2}\)
  
Elements number
40
160
360
640
No penalty
2.22
2.22
2.22
2.66
Penalty
2.59
3.18
3.16
3.24
The bold VFEM is used to indicate with the letter V the stress operator of the procedure, and FEM is the acronym for Finite Element Method
The inequalities satisfaction map is reported in Fig. 6 and compared with a standard FEM plastic strain plot.By the reported map, it can be seen that the Lagrange multipliers do respect the expected plastic hinge positions that result from classical plasticity analysis on one-dimensional models of beams. In particular, the one-dimensional positioning of the plastic hinges of the supported cantilever subjected to uniform load is at 0.58578 of the span length corresponding to 2360 mm of the beam. It can be seen that the location of the elements violating the constitutive constraints respects the expected analytical position of plastic hinge arising [9].

3.2 Holed plate

The plate with a circular hole under tension is one of the most referenced examples for calculating the influence of voids and inclusions on the resistance of structures. The present section reported the calculations of a square plate with a central hole. The typical circular hole has been addressed using the proposed limit analysis procedure with the penalty factor correction. The results have been compared with the literature. Furthermore, a square plate with an elliptical hole has been calculated for different ellipticity values.

3.3 Circular hole

The analyzed structure was a square plate with a central circular hole subjected to uniform pressure loads \(P_1\) and \(P_2\), applied to each side as shown in Fig. 7 [26].
The material properties of the calculated plates are reported in Table 2 together with the geometric data. The material is iso-resistant at yielding \(\sigma ^+=-\sigma ^-=\sigma _0\). The ratio between the hole’s radius R and the plate’s side length L is \((R/L= 0.2)\). Because of its symmetry, only a quarter of the plate has been modeled. Two load conditions have been analyzed, \(P2=0.0\) and \(P2=0.5P1\).
Table 2
Holed plate data
Young modulus E
\(2.1\cdot 10^5\) MPa
Poisson ratio \(\nu \)
0.3
Yield stress \(\sigma _0\)
\(2.0\cdot 10^2\) MPa
Side length L
10 mm
Hole radius R
2 mm
In Table 3, the calculated collapse load in terms of yield stress factor is reported for several mesh refinements and together with a step-by-step FEM result. In Fig. 8 the mesh size is represented.
Table 3
Plate with circular hole: \(s_c\) vs mesh refinement
  
\(\varvec{\mathcal {V}}FEM\) \(s_c \cdot 10^{2}\)
  
Step-by-step FEM
Elements number
64
160
280
504
64–504
\(P2=0\)
0.772
0.768
0.777
0.779
0.8080.801
\(P2=0.5P1\)
0.891
0.891
0.892
0.892
0.9230.836
The bold VFEM is used to indicate with the letter V the stress operator of the procedure, and FEM is the acronym for Finite Element Method
For the same example, the analytical solution for the structure can be found in [55]. Some numerical results are reported in [5658], and a comparison with literature results is reported in [59] where it can be seen that the calculated results with the proposed method are in agreement with those in the literature, see Table 4
The reported results have highlighted some differences between step-by-step analysis and limit design. The proposed procedure calculates the limit load through the lower bound approach, thus underestimating the limit load. The reference FEM step-by-step approach is based on the time-step integration of the equilibrium equation of the actual elastic plastic structure using predictor–corrector algorithms. The plastic zones are highlighted in Fig. 9 It can be seen that the optimization program converges to the FEM plastic zone.
In Fig. 10, a representation of the displaced elements obtained from the dislocation map as a result of the optimization program is reported. The dislocated elements can be seen that take into account the nonconforming map due to the nodal parameters \(\textbf{v}^p\). In the figure the corresponding residual stresses have been mapped that represent the Mè lan residuals.

3.4 Elliptical hole

In the following, the void inclusion in the plate has been considered to be elliptic with different aspect ratios. The radii of the ellipse, say a and b, have been varying, as shown in Fig. 11, and the collapse multiplier \(s_{c}\) has been calculated for different ellipticity \(\mathcal {E}=\frac{b-a}{b}\). The plate has been considered subject to three load conditions, \(P_2=0\), \(P_1=P_2\), and \(P_1=0.5P_2\). The results are reported in Table 5. The resulting collapse multiplier has confirmed that the plate’s resistance has diminished for flattened holes. The results depend on the stress concentration about the hole end that diverges as \(\mathcal {E} \rightarrow {1}\). The load direction influences the effect since the load perpendicular to the void’s largest direction, \(P_1\), produces a higher stress concentration than the parallel load, \(P_2\).A plastic deformation map, obtained through the proposed procedure, compared with the plastic strain at incoming collapse produced by a standard FEM analysis is shown in Fig. 13. The combined effect of both loads results in an intermediate behavior. In Fig. 12, a set of mesh type for each ellipticity of the hole is reported. In the case of ellipticity greater than 0.6, the collapse multiplier when \(P_2=0\) is less than the multiplier when \(P_2 \ne 0\).
Table 4
Present solution compared with literature results
 
Litterature results
 
Authors and method
\(P2=0\)
\(P2=0.5P1\)
Tran et al. (dual algorithm)
0.803
.0911
Zhang et al. (BEM LB)
0.789
0.907
Corradi et al. (Linear Programming Approach)
0.69
Nguyen et al (dual algorithm).
0.807
0.911
Present (\(\varvec{\mathcal {V}}FEM\))
0.779
0.892

3.5 A cylindrical pipe with inner pressure

The case of a cylinder loaded by internal pressure has been analyzed considering the generic cross-section subjected to outward internal pressure. The material of the cylinder was iso-resistant linearly-elastic perfectly-plastic, the Young modulus was \(E=2.1\cdot 10^5\)MPa, and the Poisson ratio \(\nu =0.3\); the tension and compression limit stresses were opposite: \(\sigma ^+=-\sigma ^-=300\) MPa. The geometry of the structure is an annulus whose inner radius is \(R_1= 10\) mm and whose outer radius is \(R_2=15\) mm. Due to the constitutive law of the material, the geometry, and the load’s symmetry, only a quarter of the annulus has been considered, see Fig. 14.
The limit pressure ratio with respect to the material’s limit stress has been assumed as the pipe’s collapse multiplier, \(s_c=\frac{P_c}{\sigma _0}\); it has been varied depending on the element number used for the annulus discretization. The minimum result \(s_c=4.67 \cdot 10^{-2}\) is obtained with 75 elements and no penalty factor. The maximum, corresponding to 507 elements mesh and penalty implementation, is \(s_c=5.92 \cdot 10^{-2}\). The reference value, obtained by step-by-step FEM routine, is \(s_c^\textrm{FEM}=5.57\cdot 10^{-2}\). The comparison between the results considering different meshes see Fig. 15 with or without penalty factor is reported in Table 6. From the observation of the results, it can be noticed that the procedure converges from above for increasing element numbers because the compatibility inequalities have been satisfied in mean over the elements. A graphic representation of the violation of compatibility inequality at incoming collapse is reported in Fig. 16.
This means that compatibility can be violated locally. However, the present example confirms that the linearization of the compatibility domain produces underestimated results. The introduction of the penalty factor increases the calculated value of the load’s collapse multiplier of \(0.27\%\) and reduces the error with the reference step-by-step calculation from 19 to \(6\%\).

3.6 Porous structure

The collapse load of a porous structure is analyzed considering elementary load conditions that simulate a nominal uniform uniaxial stress acting in the structure’s interior. A square portion is considered, assuming the plane stress hypothesis. It is representative of the average constituent of the material and can be assumed as the elementary volume element, RVE. Different loads have been applied to get an estimate of the limit domain of the material, starting from the detailed micro-structural description and furnishing the average limit of the stress in any direction of the element to establish a starting point for the multiscale study of the material constitutive behavior.
The element is constituted by elastic plastic material and is supposed to be isotropic in the elastic range. The constitutive parameters of the material have been deduced from the cortical tissue of the human femur bone as described in [20]. The RVE represents a trabecular zone element with volume density influenced by the presence of void inclusions. The volume fraction of the solid tissue is \(\alpha \), calculated as the ratio between the solid volume \(V_m\) and the total volume \(V_t\):
$$\begin{aligned} \alpha =\frac{V_m}{V_\textrm{tot}}=0.655 \end{aligned}$$
The elastic parameters are the Young modulus \(E=20.0 \cdot 10^3\) MPa, and Poisson ratio \(\nu =0.1\). The two different stress limits in compression and tension are the inelastic properties, \(\sigma ^-=-140\) MPa, and \(\sigma ^+=14\) MPa, respectively. Perfect plasticity is assumed in both cases.
Table 5
Plate with elliptical hole:\(s_{c}\) vs ellipticity
  
\(\varvec{\mathcal {V}}\)FEM \(s_{c}\)
 
\(\mathcal {E}\)
\(P_2=0\)
\(P_2=0.5P_1\)
\(P_2=P_1\)
0.50
0.895
1.02
0.925
0.67
0.591
0.642
0.713
0.75
0.485
0.507
0.542
0.80
0.423
0.431
0.461
The trabecular material’s RVE has been loaded by uniform pressure \(p=1.0\) MPa, on two sides as described in Fig. 17. Two load conditions have been applied to the RVE to simulate the compression and shear stresses. The obtained collapse loads have been assumed as the strength of the material under the two elementary uniaxial stress conditions. As a reference, a third example with the same RVE under compression, but assuming iso-resistant material at the limit, having \(\sigma ^+=-\sigma ^-=140\) MPa, has been considered. The limit stresses obtained from the analyses are summarized in Table 7:
The collapse load calculated for uniso-resistant material allows us to evaluate the relationship between shear and compression uniaxial yield stresses.
$$\begin{aligned} \tau _{xy}^0= 0.65\ {\sigma ^+} \end{aligned}$$
(23)
In Fig. 19, the plastic deformation and the violation of the compatibility inequality map are reported.
It has resulted greater than both Mises’, \(\tau _{xy}^0= 0.577\ \sigma ^+\), and Tresca’s, \(\tau _{xy}^0= 0.50\ \sigma ^+\) relationships and suggests the compression limit’s influence on the shear resistance that lets the coefficient of the limit tensile stress increase with respect to solid element. Indeed, the presence of voids produces the coupling of the stress state within the RVE that modifies with respect to a perfectly solid RVE, modifying the local directions of tensile and compression principal direction of the stress. The resulting stress limit for the porous volume is in good agreement with the result reported in [20] where the yield strain is \(\epsilon _y=0.0069\) for the trabecular tissue with \(\alpha =0.7\) and Young modulus \(E_t=4.0 \cdot 10^3\) MPa resulting in \(\sigma _y=27.6\) MPa.
The yield shear stress in Eq. (23) is the intercept of the plastic domain of the homogenized material with the positive \(\tau _{xy}\) axis in the stress space. Through the limit stresses for the elementary load cases, it is possible to calculate the complete intercepts with the axis of the homogenized material. Hence, one obtains the complete representation of the linearized limit domain to be used for limit analysis at the macro-scale of structures made of the material previously analyzed at the micro-scale.
Table 6
Cylindrical pipe \(s_c\) vs. mesh refinement
  
\(\varvec{\mathcal {V}}\)FEM \(s_c\cdot 10^{2}\)
   
Elements number
75
200
300
432
507
No penalty
4.67
4.67
4.67
4.67
4.67
Penalty
6.32
6.14
6.07
5.96
5.92

4 Conclusions

The work has presented the direct application of the lower bound theorem of limit analysis via finite elements employing Mèlan’s formulations for determining the collapse load of structures subjected to either monotonic or randomly variable loads. A methodological procedure is expounded that hinges upon the multiplicity of solutions to the equilibrium equations under the specified loading conditions. A comprehensive \(\varvec{\mathcal {V}}\) FEM procedure is outlined, delineating the representation of equilibrated stress’s linear variety as a composite of the linear variety spanned through \(\varvec{\mathcal {V}}\) and a particular solution postulating the structure to possess indefinite elasticity
The limit analysis of the structure is executed through Mèlan’s theorem, leveraging nodal parameters of dislocations and constrained displacements as optimization design variables. The formulation is characterized by its generality, accommodating linear-elastic perfectly-plastic materials with isotropic, anisotropic, iso-resistant, and uniso-resistant constitutive laws. The establishment of the limit domain of admissible stresses relies on the Mises–Scheicher equation, followed by the linearization of the limit locus to enhance accuracy, coupled with the introduction of a penalty coefficient for actual stress.
Validation of the method is performed across canonical geometries and constitutive materials, further extending its application to analyze plates with various hole geometries and multi-hole plates. Emphasis is placed on the necessity of estimating dissipated energy during loading and the method’s limitations concerning the control of ductility requirements. To address these concerns, a proposal is made to utilize \(\varvec{\mathcal {V}}\)FEM for upper-bound calculations of dissipated energy and local permanent displacements, showcasing the potential for future applications and developments in the field.
Table 7
Spongy RVE limit stress \(\left[ \text {MPa} \right] \)
Uniso-resistant
Uniso-resistant
Iso-resistant
Compression
Shear
Compression
19.54
12.85
158.0
The works [60, 61] show that the proposed eigenstress map, in terms of displacement and dislocation, has its natural evolution toward the application to the assessment of upper bounds of the dissipated energy and the local permanent displacements.

Declarations

Conflict of interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literatur
1.
Zurück zum Zitat ASCE: ASCE Standard ASCE/SEI 7-10 Minimum Design Loads for Buildings and Other Structures. American Society of Civil Engineers. (2010) ASCE: ASCE Standard ASCE/SEI 7-10 Minimum Design Loads for Buildings and Other Structures. American Society of Civil Engineers. (2010)
5.
Zurück zum Zitat Drucker, D.C., Greenberg, H.J., Prager, W.: The safety factor of an elastic-plastic body in plane strain. J. Appl. Mech. 18, 371–378 (1951)MathSciNetCrossRef Drucker, D.C., Greenberg, H.J., Prager, W.: The safety factor of an elastic-plastic body in plane strain. J. Appl. Mech. 18, 371–378 (1951)MathSciNetCrossRef
6.
Zurück zum Zitat Drucker, D., Prager, W., Greenberg, H.: Extended limit design theorems for continuous media. Q. Appl. Math. 9(4), 381–389 (1952)MathSciNetCrossRef Drucker, D., Prager, W., Greenberg, H.: Extended limit design theorems for continuous media. Q. Appl. Math. 9(4), 381–389 (1952)MathSciNetCrossRef
7.
Zurück zum Zitat Prager, W.: Recent developments in the mathematical theory of plasticity. J. Appl. Phys. 20, 235–241 (1949)MathSciNetCrossRef Prager, W.: Recent developments in the mathematical theory of plasticity. J. Appl. Phys. 20, 235–241 (1949)MathSciNetCrossRef
9.
Zurück zum Zitat Massonnet, C., Save, M.: Calcule Plastique des Construction. Centre Belgo-Luxembourgeois d’information de L’aceir, Bruxelles (1961) Massonnet, C., Save, M.: Calcule Plastique des Construction. Centre Belgo-Luxembourgeois d’information de L’aceir, Bruxelles (1961)
16.
Zurück zum Zitat Konig, J.A.: Shakedown of Elastic–Plastic Structures. Elsevier—Polish Scientific Publisher, Warszawa (1987) Konig, J.A.: Shakedown of Elastic–Plastic Structures. Elsevier—Polish Scientific Publisher, Warszawa (1987)
17.
Zurück zum Zitat Chen, W.F.: Limit Analysis and Soil Plasticity. Elsevier Scientific Publishing Co., Amsterdam (1975) Chen, W.F.: Limit Analysis and Soil Plasticity. Elsevier Scientific Publishing Co., Amsterdam (1975)
20.
Zurück zum Zitat Esposito, L., Minutolo, V., Gargiulo, P., Jonsson, H., Gislason, M.K., Fraldi, M.: Towards an app to estimate patient-specific perioperative femur fracture risk. Appl. Sci. (Switz.) 10(18), 6409 (2020)CrossRef Esposito, L., Minutolo, V., Gargiulo, P., Jonsson, H., Gislason, M.K., Fraldi, M.: Towards an app to estimate patient-specific perioperative femur fracture risk. Appl. Sci. (Switz.) 10(18), 6409 (2020)CrossRef
21.
Zurück zum Zitat Ricciardi, C., Jacob, D., Improta, G., Recenti, M., Gíslason, M.K., Cesarelli, G., Esposito, L., Minutolo, V., Bifulco, P., Gargiulo, P.: Improving prosthetic selection and predicting bmd from biometric measurements in patients receiving total hip arthroplasty. Diagnostics 10(10), 815 (2020)CrossRef Ricciardi, C., Jacob, D., Improta, G., Recenti, M., Gíslason, M.K., Cesarelli, G., Esposito, L., Minutolo, V., Bifulco, P., Gargiulo, P.: Improving prosthetic selection and predicting bmd from biometric measurements in patients receiving total hip arthroplasty. Diagnostics 10(10), 815 (2020)CrossRef
22.
Zurück zum Zitat Minutolo, V., Esposito, L., Sacco, E., Fraldi, M.: Designing stress for optimizing and toughening truss-like structures. Meccanica 55(8), 1603–1622 (2020)MathSciNetCrossRef Minutolo, V., Esposito, L., Sacco, E., Fraldi, M.: Designing stress for optimizing and toughening truss-like structures. Meccanica 55(8), 1603–1622 (2020)MathSciNetCrossRef
24.
Zurück zum Zitat Zienkiewicz, O., Taylor, R.: The Finite Element Method, 5th edn. Butterworth-Heinemann, Oxford (2000) Zienkiewicz, O., Taylor, R.: The Finite Element Method, 5th edn. Butterworth-Heinemann, Oxford (2000)
25.
Zurück zum Zitat Crisfield, M.A.: Non-linear Finite Element Analysis of Solids and Structures. Wiley, Chichester (1997) Crisfield, M.A.: Non-linear Finite Element Analysis of Solids and Structures. Wiley, Chichester (1997)
26.
Zurück zum Zitat Lubliner, J.: Plasticity Theory. Macmillan Pub Co, New York (1990) Lubliner, J.: Plasticity Theory. Macmillan Pub Co, New York (1990)
27.
28.
Zurück zum Zitat Simo, J.C., Hughes, T.J.R.: Computational Inelasticity. Springer, New York (1997) Simo, J.C., Hughes, T.J.R.: Computational Inelasticity. Springer, New York (1997)
35.
Zurück zum Zitat Zona, R., Esposito, L., Palladino, S., Totaro, E., Minutolo, V.: Semianalytical lower-bound limit analysis of domes and vaults. Appl. Sci. (Switz.) 12, 9155 (2022)CrossRef Zona, R., Esposito, L., Palladino, S., Totaro, E., Minutolo, V.: Semianalytical lower-bound limit analysis of domes and vaults. Appl. Sci. (Switz.) 12, 9155 (2022)CrossRef
37.
Zurück zum Zitat Vandepitte, D., Lagae, G.: Buckling of spherical domes made of microconcrete and creep buckling of such domes under long-term loading. In: Bevilacqua, L., Feijóo, R., Valid, R. (eds.) Inelastic Behaviour of Plates and Shells, pp. 291–311. Springer, Berlin (1986)CrossRef Vandepitte, D., Lagae, G.: Buckling of spherical domes made of microconcrete and creep buckling of such domes under long-term loading. In: Bevilacqua, L., Feijóo, R., Valid, R. (eds.) Inelastic Behaviour of Plates and Shells, pp. 291–311. Springer, Berlin (1986)CrossRef
38.
Zurück zum Zitat Piero, G.D.: Limit analysis and no-tension materials. Int. J. Plast. 14, 259–271 (1998)CrossRef Piero, G.D.: Limit analysis and no-tension materials. Int. J. Plast. 14, 259–271 (1998)CrossRef
39.
Zurück zum Zitat Castigliano: Theorie de l’equilibre des systemes elastiques et ses applications. Elastic stresses in structures (1919) Castigliano: Theorie de l’equilibre des systemes elastiques et ses applications. Elastic stresses in structures (1919)
40.
Zurück zum Zitat Heyman, J.: The stone skeleton. Int. J. Solids Struct. 2, 249–279 (1966)CrossRef Heyman, J.: The stone skeleton. Int. J. Solids Struct. 2, 249–279 (1966)CrossRef
41.
Zurück zum Zitat Crisfield: Finite element and mechanism methods for the analysis of masonry and brickwork arches. Transp. Res. Lab. 19 (1985) Crisfield: Finite element and mechanism methods for the analysis of masonry and brickwork arches. Transp. Res. Lab. 19 (1985)
42.
Zurück zum Zitat Melbourne, C., Gilbert, M.: Computational modelling of masonry, brickwork and blockwork structures. Model. Mason. Arch. Bridges 19 (2001) Melbourne, C., Gilbert, M.: Computational modelling of masonry, brickwork and blockwork structures. Model. Mason. Arch. Bridges 19 (2001)
60.
Zurück zum Zitat Vitiello, E.: Upper bounds to plastic strains in shake-down of structures subjected to cyclic loads. Meccanica 7(3), 205–213 (1972)CrossRef Vitiello, E.: Upper bounds to plastic strains in shake-down of structures subjected to cyclic loads. Meccanica 7(3), 205–213 (1972)CrossRef
61.
Zurück zum Zitat Palladino, S., Esposito, L., Ferla, P., Totaro, E., Zona, R., Minutolo, V.: Experimental and numerical evaluation of residual displacement and ductility in ratcheting and shakedown of an aluminum beam. Appl. Sci. (Switz.) 10(10), 3610 (2020)CrossRef Palladino, S., Esposito, L., Ferla, P., Totaro, E., Zona, R., Minutolo, V.: Experimental and numerical evaluation of residual displacement and ductility in ratcheting and shakedown of an aluminum beam. Appl. Sci. (Switz.) 10(10), 3610 (2020)CrossRef
Metadaten
Titel
A dislocation-based finite element method for plastic collapse assessment in solid mechanics
verfasst von
Renato Zona
Vincenzo Minutolo
Publikationsdatum
25.04.2024
Verlag
Springer Berlin Heidelberg
Erschienen in
Archive of Applied Mechanics
Print ISSN: 0939-1533
Elektronische ISSN: 1432-0681
DOI
https://doi.org/10.1007/s00419-024-02594-6

    Marktübersichten

    Die im Laufe eines Jahres in der „adhäsion“ veröffentlichten Marktübersichten helfen Anwendern verschiedenster Branchen, sich einen gezielten Überblick über Lieferantenangebote zu verschaffen.