DATA-DRIVEN REPRESENTATION AND CLUSTERING DISCRETIZATION METHOD AND SYSTEM FOR DESIGN OPTIMIZATION AND/OR PERFORMANCE PREDICTION OF MATERIAL SYSTEMS AND APPLICATIONS OF SAME

A method and system for design optimization and/or performance prediction of a material system includes generating a representation of the material system at a number of scales, the representation at a scale comprising microstructure volume elements (MVE) of building blocks of the material system at said scale; providing inputs to the MVEs; collecting data of response fields of the MVE computed from a material model of the material system over a predefined set of material properties and boundary conditions; applying machine learning to the collected data to generate clusters; computing an interaction tensor of interactions of each cluster with each of the other clusters; and solving an governing partial differential equation using the generated clusters and the computed interactions to result in a response prediction usable in an iterative scheme in a multiscale model for the material system. The performance of each scale can be predicted for design optimization.

Skip to: Description  ·  Claims  · Patent History  ·  Patent History
Description
CROSS-REFERENCE TO RELATED PATENT APPLICATION

This application claims priority to and the benefit of, pursuant to 35 U.S.C. § 119(e), of U.S. provisional patent application Ser. No. 62/731,381, filed Sep. 14, 2018, entitled “MULTISCALE MODELING PLATFORM AND APPLICATIONS OF SAME”, by Wing Kam Liu, Jiaying Gao, Cheng Yu and Orion L. Kafka, which is incorporated herein by reference in its entirety.

This application is related to a co-pending U.S. patent application, entitled “INTEGRATED PROCESS-STRUCTURE-PROPERTY MODELING FRAMEWORKS AND METHODS FOR DESIGN OPTIMIZATION AND/OR PERFORMANCE PREDICTION OF MATERIAL SYSTEMS AND APPLICATIONS OF SAME”, by Wing Kam Liu, Jiaying Gao, Cheng Yu, and Orion L. Kafka, with Attorney Docket No. 0116936.152US2, filed on the same day that this application is filed, and with the same assignee as that of this application, which is incorporated herein by reference in its entirety.

FIELD OF THE INVENTION

The invention relates generally to materials, and more particularly, to a specific method and system to aggregate dissimilar material geometry, properties, and interactions to predict combined properties and performance and applications of the same.

BACKGROUND OF THE INVENTION

The background description provided herein is for the purpose of generally presenting the context of the invention. The subject matter discussed in the background of the invention section should not be assumed to be prior art merely as a result of its mention in the background of the invention section. Similarly, a problem mentioned in the background of the invention section or associated with the subject matter of the background of the invention section should not be assumed to have been previously recognized in the prior art. The subject matter in the background of the invention section merely represents different approaches, which in and of themselves may also be inventions. Work of the presently named inventors, to the extent it is described in the background of the invention section, as well as aspects of the description that may not otherwise qualify as prior art at the time of filing, are neither expressly nor impliedly admitted as prior art against the invention.

Microstructured materials are aggregates of individual components. Although many approximations exist, high-fidelity modeling of a complex material typically involves building a numerical model of its microstructure. In the field of computational mechanics, virtual tests for the microstructure responses are often carried out using high-fidelity numerical methods, such as the Finite Element Method (FEM), Finite Volume Method (FVM), Fast Fourier Transform (FFT), and mesh free. This process usually requires building an explicit mesh in software for the microstructure, where the mesh is the virtual representation of the microstructure. This is typically a spatial decomposition where the elements that make up the discrete units of the decomposition follow specific rules (e.g., non-negative Jacobian for FEM, fixed spacing grid in FFT). Although there are differences, these spatial discretizations are simply referred as a “meshes.” In order to capture all the details of the microstructure, the mesh usually requires very fine resolution. This discretization of the material based on space is similar to the concept of a Riemann integral, as illustrated in FIG. 2. Typically, a fine mesh of very simple fiber composite materials may contain millions of elements.

Once the mesh for a microstructure is constructed, loads can be applied to the mesh to perform virtual testing. Various information can be obtained from this virtual testing, such as elastic properties of the material, effective strength of the microstructure, and local stress distributions to study possible microstructure failure. However, resolving a fine mesh requires significant computational resource, such as High Performance Cluster computing or GPU computing, and a long execution time for the finite element software.

For a Microstructure Volume Element (a representation of the complex material microstructure), or MVE in short, at any given scale, the local and homogenized behavior at the current scale might be computed as a function of the homogenized behavior at subscales using traditional numerical methods, such as finite element analysis. However, the computation time is proportional to the total number of elements across all scales. Each added scale compounds the difficulty. Therefore, the computation time can be high for even a simple 2-scale model if the MVE contains many elements—a typical MVE might contain O(1e6) elements, and this would have to be solved for each element at the larger scale. This means cost is very high. When done with finite elements for two scales, this is sometimes called an “FE2”. Because of the cost, there is limited practical application of such schemes, and they have seen little use.

Therefore, a heretofore unaddressed need exists in the art to address the aforementioned deficiencies and inadequacies.

SUMMARY OF THE INVENTION

In order to counter the time-consuming microstructure simulation, one of the objectives of this invention is to develop a data-driven domain decomposition approach that is suitable to accelerating the numerical simulation of the microstructure responses.

In one aspect, the invention relates to a method for design optimization and/or performance prediction of a material system. In one embodiment, the method includes generating a representation of the material system at a number of scales, wherein the representation at a scale comprises microstructure volume elements (MVE) that are building blocks of the material system at said scale; collecting data of response fields of the MVE computed from a material model of the material system over a predefined set of material properties and boundary conditions; applying machine learning to the collected data of response fields to generate clusters that minimize a distance between points in a nominal response space within each cluster; computing an interaction tensor of interactions of each cluster with each of the other clusters; manipulating the governing partial differential equation (PDE) using Green's function to form a generalized Lippmann-Schwinger integral equation; and solving the integral equation using the generated clusters and the computed interactions to result in a response prediction that is usable for the design optimization and/or performance prediction of the material system.

In one embodiment, the method further comprises passing the resulted response prediction to a next coarser scale as an overall response of that building block, and iterating the process until a final scale is reached.

In one embodiment, the building blocks are defined by material properties and structural descriptors obtained by modeling or experimental observations and encoded in a domain decomposition of structures for identifying locations and properties of each phase within the building blocks.

In one embodiment, the structural descriptors comprise characteristic length and geometry.

In one embodiment, the boundary conditions are chosen to satisfy the Hill-Mandel condition.

In one embodiment, the collected data of response fields comprise a strain concentration tensor, a deformation concentration tensor, stress tensor (e.g., PK-I stress, Cauchy stress), plastic strain tensor, thermal gradient, or the like.

In one embodiment, the machine learning comprises unsupervised machine learning and/or supervised machine learning.

In one embodiment, the machine learning is performed with a self-organizing mapping (SOM) method, a k-means clustering method, or the like.

In one embodiment, the clusters are generated by marking all material points that have the same response field within the representation of the material system with a unique ID and grouping material points with the same ID.

In one embodiment, the generated clusters are a reduced representation of the material system, which reduces the number of degrees of freedom required to represent the material system.

In one embodiment, the generated clusters are a reduced order MVE of the material system.

In one embodiment, the computed interaction tensor is for all pairs of the clusters.

In one embodiment, said computing of the interaction tensor is performed with fast Fourier transform (FFT), numerical integration, or finite element method (FEM).

In one embodiment, the PDE is reformulated as a Lippmann-Schwinger equation. In one embodiment, said solving of the PDE is performed with arbitrary boundary conditions and/or material properties.

In one embodiment, the collected data of response fields, the generated clusters, and/or the computed interaction tensor are saved in one or more material system databases.

In one embodiment, said solving of the PDE is performed in real-time by accessing the one or more material system databases for the generated clusters and the computed interactions.

In another aspect, the invention relates to a method for design optimization and/or performance prediction of a material system. In one embodiment, the method includes performing an offline data compression, wherein original MVE of building blocks of the material system are compressed into clusters, and an interaction tensor of interactions of each cluster with each of the other clusters is computed; and solving an governing PDE using the clusters and the computed interactions to result in a response prediction that is usable for the design optimization and/or performance prediction of the material system.

In one embodiment, the method further includes passing the resulting response prediction to a next coarser scale as an overall response of that building block, and iterating the process until a final scale is reached.

In one embodiment, the building blocks are defined by material properties and structural descriptors obtained by modeling or experimental observations and encoded in a domain decomposition of structures for identifying locations and properties of each phase within the building blocks.

In one embodiment, the structural descriptors comprise characteristic length and geometry.

In one embodiment, when more than one scale is involved with the reduced order response prediction, the method is named Multiresolution Clustering Analysis (MCA).

In one embodiment, the boundary conditions are chosen to satisfy the Hill-Mandel condition.

In one embodiment, said performing the offline data compression comprises collecting data of response fields of the MVE computed from a material model of the material system over a predefined set of material properties and boundary conditions; applying machine learning to the collected data of response fields to generate clusters that minimize a distance between points in a nominal response space within each cluster; and computing the interaction tensor is for all pairs of the clusters.

In one embodiment, the collected data of response fields comprise a strain concentration tensor, a deformation concentration tensor, stress tensor (e.g., PK-I stress, Cauchy stress), plastic strain tensor, thermal gradient, or the like.

In one embodiment, the machine learning comprises unsupervised machine learning and/or supervised machine learning.

In one embodiment, the machine learning is performed with an SOM method, a k-means clustering method, or the like.

In one embodiment, the clusters are generated by marking all material points having the same response field within the representation of the material system with a unique ID and grouping material points with the same ID.

In one embodiment, the clusters are a reduced representation of the material system, which reduces the number of degrees of freedom required to represent the material system.

In one embodiment, the clusters are a reduced order MVE of the material system.

In one embodiment, said computing the interaction tensor is performed with FFT, numerical integration, or FEM.

In one embodiment, the PDE is a Lippmann-Schwinger equation. In one embodiment, said solving the PDE is performed with arbitrary boundary conditions and material properties.

In one embodiment, the collected data of response fields, the generated clusters, and/or the computed interaction tensor are saved in one or more material system databases.

In one embodiment, said solving the PDE is performed with online accessing the one or more material system databases for the generated clusters and the computed interactions.

In yet another aspect, the invention relates to a non-transitory tangible computer-readable medium storing instructions which, when executed by one or more processors, cause a system to perform the above-disclosed method for design optimization and/or performance prediction of a material system.

In a further aspect, the invention relates to a computational system for design optimization and/or performance prediction of a material system. In one embodiment, the computational system includes one or more computing devices comprising one or more processors; and a non-transitory tangible computer-readable medium storing instructions which, when executed by the one or more processors, cause the one or more computing devices to perform the above-disclosed method for design optimization and/or performance prediction of a material system.

In one aspect, the invention relates to a material system database usable for conducting efficient and accurate multiscale modeling of a material system, In one embodiment, the material system database includes clusters for a plurality of material systems, each of which groups all material points having a same response field within MVE of a respective material system with a unique ID; interaction tensors, each of which represents interactions of all pairs of the clusters for the respective material system; and response predictions computed based on the clusters and the interaction tensors.

In one embodiment, the clusters are generated by applying machine learning to data of response fields of the MVE computed from a material model of the respective material system over a predefined set of material properties and boundary conditions.

In one embodiment, the interaction tensors are computed with FFT, numerical integration, or FEM

In one embodiment, the responses predictions are obtained by solving a governing PDE using the clusters and the computed interactions. In one embodiment, the responses predictions comprise at least effective stiffness, yield strength, thermal conductivity, damage initiation, and FIP.

In one embodiment, the material system database is configured such that some of the response predictions are assigned as a training set for training a different machine learning that connects processes/structures to responses/properties of the material system directly without going through the clustering and interaction computing processes at all; and some or all of the remaining response predictions are assigned as a validation set for validating the efficiency and accuracy of the multiscale modeling of the material system.

In one embodiment, the material system database is generated with predictive reduced order models. In one embodiment, the predictive reduced order models comprise a self-consistent clustering analysis (SCA) model, a virtual clustering analysis (VCA) model, and/or an FEM clustering analysis (FCA) model.

In one embodiment, the material system database is updatable, editable, accessible, and searchable.

In another aspect, the invention relates to a method of applying the above-disclosed material system database for design optimization and/or performance prediction of a material system. In one embodiment, the method includes training a neural network with data of the material system database; and predicting real-time responses during a loading process performed using the trained neueral network, wherein the real-time responses are used for the design optimization and/or performance prediction of a material system.

In one embodiment, the method further includes performing a topology optimization to design a structure with microstructure information.

In one embodiment, the neural network comprises a feed forward neural network (FFNN) and/or a convolutional neural network (CNN).

In yet another aspect, the invention relates to a non-transitory tangible computer-readable medium storing instructions which, when executed by one or more processors, cause a system to perform the above method for design optimization and/or performance prediction of a material system.

These and other aspects of the invention will become apparent from the following description of the preferred embodiment taken in conjunction with the following drawings, although variations and modifications therein may be affected without departing from the spirit and scope of the novel concepts of the invention.

BRIEF DESCRIPTION OF THE DRAWINGS

The following drawings form part of the present specification and are included to further demonstrate certain aspects of the invention. The invention may be better understood by reference to one or more of these drawings in combination with the detailed description of specific embodiments presented herein. The drawings described below are for illustration purposes only. The drawings are not intended to limit the scope of the present teachings in any way.

FIG. 1 illustrates schematically an overall flow of data through the system according to embodiments of the invention. Key: heavy bordered boxes are operations, typically including computer code; these act upon various data contained within the parallelograms and typically stored on disk or in memory depending upon how the operations are embodied. In the more complex operations, information or steps commonly used to produce a desired operation are listed; these appear as numbered, round-cornered boxes. The system starts with a representation of the material system at a finite number of scales. Any particular scale is composed of fundamental building blocks, the size of which is defined by a characteristic length. The composition of the building block is defined by modeling or experimental observations. These are then encoded in a detailed spatial decomposition of the structure (sometimes called a “mesh”), used to describe the location and properties of each phase within the building block. Generation of this high-resolution description is termed microstructure generation and may be relevant to any scale. A predefined set (usually carefully selected and simplified) of material properties and boundary conditions are supplied to a direct numerical solver to compute nominal response fields. Unsupervised machine learning is applied to these fields to generate clusters that minimize the distance between points in nominal response space within each cluster. This produces a reduced representation of the material building block. The interaction between each cluster (or, the influence that a unit load applied to one cluster has on other clusters) can be computed and stored. Using precomputed clusters and interactions, the solution to the relevant governing partial differential equation (PDE) with arbitrary boundary conditions and material properties (not necessarily the simplified ones used to compute clusters) is fast in the response prediction stage. The results of the response prediction are passed to the next coarser scale, as the overall response of that building block, and the process thus proceeds.

FIG. 2 shows discretization of partial differential equations through, panel (a), domain-based decomposition and panel (b), response-based decomposition clustering e.g., with unsupervised learning, according to embodiments of the invention.

FIG. 3 shows example material systems with two or more relevant length scales that could be modeled by the method and system, according to embodiments of the invention.

FIG. 4 shows an MVE model of a unidirectional carbon fiber reinforced composite containing 36 million voxels, according to embodiments of the invention. On the left, the blue cylinders represent carbon fiber, and the red is a polymetric epoxy resin matrix materials. The overall (homogenized) response to an applied load in the y-direction is plotted in the top right, and the local stress contours are plotted on the bottom right. The direct numerical simulation (DNS) process to compute this single load history takes about 200 hours using 80 computer processors.

FIG. 5 is an illustration of the 2-scale multiscale problem and its complexity using traditional finite element method (top panel), where at the macroscale, a mesh of a generic part subject to some boundary condition is shown and at each material point within these elements, a microscale response corresponding to the behavior of a generic subscale, shown in this example with spherical inclusions, is computed; and an illustration of two-scale multiscale clustering method and its capability in order reduction (bottom panel), according to embodiments of the invention. In this illustration, each color represents order 1 degrees of freedom, rather than each blue-bounded region within the top illustration.

FIG. 6 shows three-step “training” process, as one possible way to compute the clusters and interaction tensor for macroscale geometry (top panel) and microscale geometry (bottom panel), according to embodiments of the invention.

FIG. 7 shows a three-scale FRP modeling framework, according to embodiments of the invention. The macroscale model is a woven laminate composite model, built as a Finite Element mesh. Each integration point in the macroscale model is represented by a woven MVE, which can have different tow size, tow spacing, and tow angle (the one shown is 90° tow angle). Each integration point in the mesoscale model is represented by a UD MVE, which can have different fiber orientation, fiber volume fraction, and matrix-fiber interfacial strength. If all three scales are discretized with FE meshes, the total DOF for this full-field system could be 3.3×1015. Using the present method and replace mesoscale and microscale with ROMs, the total DOF is reduced down to 1.6×106, a reduction of nine orders of magnitude.

FIG. 8 shows MVEs generated from a number of sources, according to embodiments of the invention. This example for metallic materials shows grains measured using x-ray diffraction, reconstructed from a statistical description, and predicted from a processing model. Defects, voids in this case, can be from measurements or predictions, here two experimental methods (x-ray tomography and FIB-SEM serial sectioning) are given as examples. These input data are paired (spatially), and used for the MVE, resulting in response predictions based directly on experimental images.

FIG. 9 shows a multiscale cluster-based process according to embodiments of the invention. Left: the original DNS description of a UD MVE. Center: cluster-based description (showing 2 clusters in the fiber phase, 8 in the matrix), visualized on the underlying voxel mesh. Right: numerical homogenization results for overall stress/strain. The DNS reference solution is shown in red with pointwise 5% error bars. Cf the reduced order solutions with 8 clusters (blue) and 16 clusters (green).

FIG. 10 shows a physics guide NN may be “layered” on top of the MVE ROM according to embodiments of the invention: the NN is trained on a large database of rapidly-computed behavior. Once trained, this NN is thought to contain microstructural information similar to the ROM, but is much faster to evaluate. This alternative makes it practical to conduct microstructure-based structural optimization and design.

FIG. 11 shows an application example: a composite design, according to embodiments of the invention. The composite structure can be designed with different microscale structure (e.g., fiber shape) and mesoscale structure (e.g., fiber orientation in each ply, fiber orientation and fiber shape in each ply, as well as different weave pattern for woven composite). Key performance indices (e.g., strength and maximum deformation under external load) for the composite structure can be predicted. If those indices do not meet desired criterion, optimization routine will be called to update microscale and mesoscale structure in order to improve performance indices.

FIG. 12 shows woven composite MVE domain decomposition according to embodiments of the invention.

FIG. 13 shows a random grain structure with 35 grains, where each grain is considered a different material phase and grain-by-grain stress predictions, with progressively more clusters per grain according to embodiments of the invention, as discussed in EXAMPLE 5.

FIG. 14 is an illustration of an implementation with hierarchical modeling between mesoscale and microscale, and concurrent modeling between macroscale and mesoscale, according to embodiments of the invention. The UD MVE elastic properties are pre-computed and passed to yarn, constituting to a hierarchical modeling process. When the woven MVE is under external loading, its response is computed using both yarn (given by the UD MVE) and matrix properties. When the FE model is under external loading, its local responses is computed using the woven MVE. The FE model and woven MVE responses are computed in a concurrent fashion, establishing a concurrent modeling scheme. A combined hierarchical and concurrent modeling is thus implemented.

FIG. 15 shows stress vs. strain curves for six loading directions (woven is isotropic in the xy plane, thus the yy response is same as the xx response, and the xz response is same as the yz response, panel (a)), as discussed in EXAMPLE 8; and 3D yield surface visualized with stress states (magenta asterisks, panel (b)), according to embodiments of the invention. All asterisks represent a stress state that does not cause yielding of the material. The present method allows one to build yield surfaces for various microstructure and generate material constitutive information with minimum effort (around one minute using a personal computer). A large woven composite response database can be built to assist design of woven composite against yielding. Given a priori information on maximum service loads, the database will provide all possible woven microstructure (e.g., yarn geometry and yarn angle) and material constituents (e.g., matrix properties and yarn properties) that would prevent yielding to occur.

FIG. 16 shows offline and online stages of two-stage clustering analysis methods according to embodiments of the invention. The offline stage contains three steps as shown in the figure, which will generate a compressed RVE database. The compressed RVE can be then used to predict mechanical responses of the RVE.

FIG. 17 shows geometry for example problem: a two-phase periodic composite represented in 2D with plane strain where blue is a continuous matrix phase and yellow represents inclusions.

FIG. 18 shows color contours showing clusters distribution in the inclusions (panel a) and the matrix (panel b), according to embodiments of the invention. Note that clusters need not be spatially connected.

FIG. 19 shows component-wise magnitude plots for a) DSCAIJ, b) DVCAIJ, c) BFCAIJ, according to embodiments of the invention. Spikes along the diagonal direction for all three interaction tensor surface plots suggest self interaction has more contribution than the rest of clusters in cluster-wise stress increment. DSCAIJ and DVCAIJ have the similar magnitude along their diagonal direction due to the homogeneous reference material assumption. BFCAIJ has different magnitudes for matrix and inclusion phase along the diagonal direction, implicitly representing a heterogeneous reference material.

FIG. 20 shows plots for a) DSCAIJ, b) DVCAIJ, c) BFCAIJ in profile, according to embodiments of the invention. Note that for FCA the two regions correspond to different physical domains (matrix and inclusion).

FIG. 21 shows plots of σxxM vs. εxxM, according to embodiments of the invention. All three methods performed well, with predictions laying within 5% of the reference solution.

FIG. 22 shows plots of σyyM vs. εxxM, according to embodiments of the invention. SCA has the best agreement with the DNS solution in this case, whereas VCA and FCA deviate from the DNS solution. The causes for such deviation and improvements in VCA and FCA will be explored in the future.

FIG. 23 shows random samples of strain state, according to embodiments of the invention; two hundred final states were selected, and four evenly spaced intermediate steps to reach the final states were recorded for a total of NT=1,000 samples. All strain states will be applied to the RVE to generate corresponding stress states.

FIG. 24 is illustration of an FFNN network with one hidden layer for a linear elastic example (panel a), the collective function of the weights and biases connecting the input layer (green), the hidden layer (blue), and the output layer (red) is that of Young's Modulus E, and a stress-strain diagram (panel b), showing how the input strain is interpreted by the FFNN for a linear elastic case using linear activation functions and zero biases.

FIG. 25 is illustration of an FFNN with multiple hidden layers; NL: index of layers, NN(l): number of neurons in layer l. The formulation of the FFNN is given in Eq. (1-21) with associated interpretation of the FFNN structure. The indices i and j representing neuron id in previous layer and current layer, e.g., W12l=2 is the weight between neuron 1 in layer l=1 and neuron 2 in layer l=2.

FIG. 26 shows a histogram of the difference between the FFNN σM predictions and validation data set for all NV=150 samples according to embodiments of the invention. Most of the predictions made by FFNN has an l2 norm less than 2×10−5, showing that the FFNN produce an accurate prediction of the stress state.

FIG. 27 shows a cross-correlation between SCA and FFNN for x-, y-, and shear-directions according to embodiments of the invention. The solid black lines are the ground truth shows all perfect predictions should lay on those lines. The FFNN-predicted stress components lay around the solid black lines, and all three cases have correlation coefficients of 1, exhibiting very strong predictive confidence.

FIG. 28 shows a stress-strain plot with both the FFNN and SCA solutions for the validation data, showing that the FFNN results successfully reproduced SCA results, according to embodiments of the invention.

FIG. 29 shows an illustration of one dimensional convolutional neural network with the following setup: padding, convolution, pooling, and a feed forward neural network for regression analysis, according to embodiments of the invention. The first three steps may be repeated.

FIG. 30 shows an illustration of two dimensional convolutional neural network, according to embodiments of the invention. In the training part, the number of threshold η*=5. After 4 repeats of convolution, the data will be passed to FFNN. The dimension of input for FFNN becomes Nfl=74. The hidden layer in FFNN has 74 neurons. Finally, FFNN gives three macro strain components.

FIG. 31 shows a histogram of the l2 norm for the CNN prediction using validation data points, according to embodiments of the invention. Most of the predictions made by CNN has an l2 norm less than 1×10−5, showing the CNN produce an accurate prediction of the strain state. The l2 norm illustrate the CNN network is able to make a proper prediction of the validation data.

FIG. 32 shows a loading prediction by CNN vs. loading applied in SCA using validation data points, according to embodiments of the invention. The solid black lines are the ground truth shows all perfect predictions should lay on those lines. All three cases have correlation coefficients higher than 0.99, suggesting the trained CNN can provide a good accuracy in predicting applied strain.

FIG. 33 shows a topology optimization setup with FFNN, according to embodiments of the invention. The FFNN will be used to compute non-linear material responses to drive a new design. This replaces the constitutive law commonly used for the macroscale with a homogenized response of the microstructure for each point in the macroscale. Mathematically, σM (XM)=β(XM)FFNN(ε(XM)), ∀XM M, as defined in Eq. (1-36).

FIG. 34 shows an optimized beam structures with elastic material responses (pane a) and non-linear material responses (panel b), according to embodiments of the invention. Comparing panels a) and b), the joint of all truss members for the non-linear case is located towards the bottom side of the truss structure. This means the material non-linear responses plays an important role in determining optimized truss structure in this case.

FIG. 35 shows a topology optimization setup with FFNN and CNN, according to embodiments of the invention. For each material point within the design zone, the FFNN is used to compute the material response, be it linear or non-linear, considering the effect of microstructure. The CNN is used to incorporate microstructure damage, which will drive the optimization algorithm for a new design compared to a topology optimization with only linear material. Mathematically, this is ∀XM M (XM): σ(X)=FFNNmicroM(XM)), ∀X∈(X): d(XM)=(CNNclassify(σ(X))), as defined in Eq. (37). If any d(X) is marked as damaged in the microstructure, the XM point in the design zone containing that microstructure will be marked as damaged.

FIG. 36 shows optimized beam structures without damage constraints (panel a) and with damage constraints (panel b), according to embodiments of the invention. In panel (b) there are more truss members to avoid local stresses concentration that results in damage, as discussed in EXAMPLE 1.

FIG. 37 shows ductile materials' microstructures discretized using voxel meshes with matrix shown in blue and inclusions in red: (panel a) two-dimensional microstructure, (panel b) inside view of a three-dimensional microstructure with a fragmented inclusion surrounded by a debonding void shown in light gray, according to embodiments of the invention

FIG. 38 shows ductile materials' microstructures, according to embodiments of the invention: (panel a) two-dimensional microstructure discretized using 8 clusters, (panel b) same two-dimensional microstructure discretized using 65 clusters, and (panel c) three-dimensional microstructure discretized using 217 clusters showing two clusters in the matrix phase (two shades of blue), one cluster in the inclusion phase (red), and one cluster in the void phase (light gray).

FIG. 39 shows ductile composite with particle volume fraction of 20% (panel a); SCA converges fast to the FFT reference for the overall mechanical response of the RVE under unidirectional tension along the x direction (panel b); (panel c) a closer look at the dashed rectangle area in panel (b) shows accurate prediction of SCA with only 16 clusters; and (panel d) a CPU time saving of a factor of more than 103 achieved with SCA for the relatively well converged number of clusters, compared with the FFT reference with a 100×100×100 voxel mesh, according to embodiments of the invention.

FIG. 40 shows (panel a) SCA is fairly close to the FFT reference for the overall mechanical response of the porous RVE under uniaxial tension along the x direction; (panel b) a closer look at dashed rectangle area in (panel a) showing the relatively slow convergence; (panel c) a CPU time saving of a factor of more than 103 is achieved with SCA for relatively well-converged numbers of clusters, compared with the FFT reference with a 100×100×100 voxel mesh, according to embodiments of the invention.

FIG. 41 shows clusters (shades of red) within the debonded inclusion showing (panel a) before fragmentation, and (panel b) after fragmentation with the missing cluster turned into void, according to embodiments of the invention. Drawing direction is horizontal.

FIG. 42 shows cold drawing and FIP computation results showing the particle fragments in red, the voids in light gray, according to embodiments of the invention: (panel a) the equivalent plastic strain in the matrix in undeformed con guration, (panel b) the equivalent plastic strain in the matrix in deformed con guration after displacement reconstruction, (c) the FIP in the matrix on a new voxel mesh of the deformed con guration after mesh transfer. Note that the three figures are not in the same spatial scale, as section height in (panels b and c) is reduced by 45% compared to (panel a).

FIG. 43 shows (panel a) ductile composite with particle volume fraction of 20%; (b) SCA converges fast to the FFT reference for the overall mechanical response of the RVE under unidirectional tension along the x direction; (panel c) a closer look at the dashed rectangle area in (panel b) shows accurate prediction of SCA with only 16 clusters; (panel d) a CPU time saving of a factor of more than 103 is achieved with SCA for the relatively well-converged number of clusters, compared with the FFT reference with a 100×100×100 voxel mesh, according to embodiments of the invention.

FIG. 44 illustrates an ICME framework according to embodiments of the invention.

FIG. 45 shows epoxy matrix according to embodiments of the invention: (panel a) tension stress versus strain curve and (panel b) compression stress-strain curve.

FIG. 46 shows UD CFRP plaque (approximately 300 mm×300 mm) with fibers oriented in the vertical direction, and the tabbed specimens after waterjet cutting, according to embodiments of the invention.

FIG. 47 shows (panel a) field of view with speckle pattern on specimen with 1=120 mm, w=12.7 mm; and (panel b) region of Interest (colored area) for data analysis, according to embodiments of the invention.

FIG. 48 shows UD CFRP hat-section sample according to embodiments of the invention: (panel a) geometry; (panel b) sample with back plate; and (panel c) dynamic 3 point bending test setup.

FIG. 49 shows UD CFRP coupon FE model setup according to embodiments of the invention. In this setup, the material response at each integration point (marked as black cross) is computed using a UD RVE, according to embodiments of the invention.

FIG. 50 shows four candidate microstructure setups in the database for the UD coupon specimen according to embodiments of the invention. The database is designed so that users can assign various microstructure to the macroscale model based on their needs. “els” in this figure stands for voxel elements in the RVE, according to embodiments of the invention.

FIG. 51 shows a cross-section of UD CFRP under microscope, with magnified view shown on the right, according to embodiments of the invention.

FIG. 52 shows an algorithm flow chart for generating UD RVE using Monte Carlo method, according to embodiments of the invention.

FIG. 53 shows UD RVEs generated by Monte Carlo method, according to embodiments of the invention.

FIG. 54 shows UD RVE clustering process for setup 1 and 2 from FIG. 50, according to embodiments of the invention.

FIG. 55 shows transverse tension results of DNS and ROMs, according to embodiments of the invention. The DNS result (computed from FEM) has an error bar representing the 95% interval. The results of ROMs computed from SCA are within the 95% interval.

FIG. 56 shows UD CFRP coupon off-axial tensile simulation setup, according to embodiments of the invention.

FIG. 57 shows UD CFRP dynamic 3-point bending model setup, according to embodiments of the invention.

FIG. 58 shows UD coupon normal stress vs. normal strain, according to embodiments of the invention. The comparison shows the prediction is in a good match with the test data. The difference between prediction and experimental result is potentially caused by the microstructure variation in the real UD material. As stated in EXAMPLE 4, all material laws are strictly from experiments, without any calibrated parameters.

FIG. 59 shows contour of(panel a) Y displacement and (panel b) Y strain field, according to embodiments of the invention. The applied displacement on prediction and DIC is 0.9031 mm. In the displacement plots a), two black arrows measure the vertical distance between fringes from −0.250 mm and −0.700 mm and the difference is 3.95%. In the gray scale strain contour b), the predicted strain field is comparable to the DIC one. The difference the predictions and the DIC images is caused by microstructure variations in the real UD CFRP material, which can cause strain concentration in the real sample.

FIG. 60 shows the coupon crack formation of (panel a) numerical prediction and (panel b) experimental result, according to embodiments of the invention. Since the numerical model assumes perfect materials without microstructure variations, the predicted pattern deviates from the test result

FIG. 61 shows a magnified view of coupon local microstructure for marked macroscale elements at a) dy=1.40 mm; b) dy=1.41 mm; c) dy=1.42 mm and local UD RVEs at d) dy=1.40 mm; e) dy=1.41 mm; f) dy=1.42 mm, according to embodiments of the invention. All RVEs are shown in undeformed configuration.

FIG. 62 shows an UD hat-section after the impact from (panel a) prediction (panel b) experiment, according to embodiments of the invention. The damaged zones on the sidewalls are marked by yellow ellipses. It can be observed that the hat-section is being pushed inwards upon the impact, and the impactor will cause a dent on the hat-section.

FIG. 63 shows UD hat-section mesh with magnified view of three columns of elements in the through-thickness direction, according to embodiments of the invention. The impactor is hidden for clarity. This setup shows the location of three columns of elements (12 elements per column, representing all 12 layers of UD laminae) across the thickness direction in the UD hat-section mesh. Each element is represented by one UD RVE, whose fiber orientation matches with the UD layup given. All UD RVEs are color-coded according to the fiber orientation.

FIG. 64 shows an isometric view of the hat-section von Mises stress contour at (panel a) Upon impact; (panel b) dz=4.85 mm; and (panel c) dz=6.85 mm, according to embodiments of the invention. The impactor is hidden for clarity. The present ICME framework is capable of visualizing structural level responses and the microstructure responses. In this case, only those elements mentioned in FIG. 63 are visualized.

FIG. 65 shows a magnified view of the hat-section with UD CFRP microstructure evolution at (panel a) dz=5.91 mm; (panel b) dz=6.13 mm; and (panel c) dz=6.78 mm, according to embodiments of the invention. The UD microstructure damage processes in two marked elements are visualized.

FIG. 66 shows an SCA flow chart for the solving the Lippmann-Schwinger equation, according to embodiments of the invention.

FIG. 67 shows clustering results of the multi-inclusion system based on the elastic strain concentration tensor A(x), according to embodiments of the invention. The numbers of clusters in the matrix and inclusion are denoted by km and ki, respectively.

FIG. 68 shows stress-strain curves under uniaxial tension and pure shear loading conditions predicted by the regression-based and projection based self-consistent scheme according to embodiments of the invention. The solid lines represent the DNS results for comparison.

FIG. 69 shows schematic of the simple 3D void geometry according to embodiments of the invention.

FIG. 70 shows RVE cut in half to show the (panel a) elastic region solution and (panel b) resulting clusters around the inclusion; (panel c) the solution at the onset of plasticity and (panel d) the resulting clusters; and (panel e) after plasticity fully develops, (panel f) the resulting clusters, according to embodiments of the invention.

FIG. 71 shows an overall stress-strain response of the single inclusion, based on different choices of cluster, according to embodiments of the invention.

FIG. 72 shows RVE including 35 equiaxed, randomly oriented grains (as shown by the inverse pole figure color map) with (a) 20×20×20 and (b) 40×40×40 voxel mesh, according to embodiments of the invention.

FIG. 73 shows σ33 versus 33 using CPFEM and CPSCA respectively, showing convergence with mesh size and number of clusters, according to embodiments of the invention.

FIG. 74 shows volume plots of S33 for six different cases: (panel a) the 203 mesh with CPFEM, (panel b) the 303 mesh with CPFEM, (panel c) the 403 mesh with CPFEM clusters, (panel d) the 403 mesh with 35 clusters, (panel e) the 403 with 70 clusters, and (panel f) the 403 mesh with 140 clusters, according to embodiments of the invention. The 35 cluster case has one cluster per grain, whereas the 140 cluster case has four clusters per grain. Opacity scales with stress level.

FIG. 75 shows Temperature dependence of horizontal shift factor for unfilled and filled rubber where aT, T and T0 stand for horizontal shift factor, temperature, and reference temperature respectively according to embodiments of the invention.

FIG. 76 shows measured shear storage modulus of unfilled and filled rubbers according to embodiments of the invention.

FIG. 77 shows measured shear loss modulus of unfilled and filled rubbers according to embodiments of the invention.

FIG. 78 shows measured loss tangent of filled rubber according to embodiments of the invention. The peak of the filled rubber is less than that of unfilled rubber. The loss tangent in the low frequency region is higher than that of unfilled rubber.

FIG. shows rubber matrix material hidden according to embodiments of the invention.

FIG. 80 shows predicted Tan(δ) of filled rubber compared to experimental results according to embodiments of the invention.

FIG. 81 shows (panel a) shear storage moduli G′ and (panel b) shear loss moduli G″ comparison between FFT and experimental results according to embodiments of the invention.

FIG. 82 shows Offline and Online Stages for Filled Rubber according to embodiments of the invention.

FIG. 83 shows clusters of filled rubber in matrix phase and filler phase according to embodiments of the invention. Domain decomposition of 2-phase filled rubber into reduced order model represented by 64 clusters. Left shows Original voxel mesh for the 2-phase filled rubber; Right shows Compressed 2-phase filled rubber model with 32 clusters in the matrix phase and 32 clusters in the filler phase.

FIG. 84 shows Tan(δ) of filled rubber computed by SCA according to embodiments of the invention.

FIG. 85 shows (panel a) shear storage moduli G′ plots and (panel b) shear soss moduli G″ plots, according to embodiments of the invention

FIG. 86 shows filled rubber with an interphase of 9.74 nm and clusters of filled rubber in matrix phase, filler phase, and interphase, according to embodiments of the invention. Domain decomposition of 3-phase filled rubber into reduced order model represented by 96 clusters. Left: Original voxel mesh for the 3-phase filled rubber; Right: Compressed 3-phase filled rubber model with 32 clusters in the matrix phase, 32 clusters in the interphase phase and 32 clusters in the filler phase.

FIG. 87 shows predicted master curve of filled rubber by FFT and SCA vs. Experimental Results according to embodiments of the invention.

FIG. 88 shows (panel a) Predicted G′ and (panel b) Predicted G″ of Filled Rubber vs. Experimental Results according to embodiments of the invention.

FIG. 89 shows Overall diagram of the computational scheme according to embodiments of the invention. Geometry, build process parameters, material, and loading conditions must be specified. These are used to conduct a thermal analysis and a macroscale stress analysis. For each material point X within these two models, an element-wise sub-model is constructed to represent a possible state at that point. This uses local thermal history and strain history to determine the microstructure (void geometry) and deformation history, respectively. These are used to predict the microscale evolution of state variables such as plasticity and damage, which are homogenized (e.g., by taking the l norm of the domain) and used as element-wise estimators of part-level susceptibility to failure.

FIG. 90 shows diagram of the defect estimation and database building process according to embodiments of the invention. In the first part, a relationship between solidification cooling rate (SCR) and void volume fraction (Vf) is determined using process modeling and X-ray computed tomography. Second, 2(a) the subsets of the images acquired with X-ray tomography are selected on the basis of Vf, such that the expected range of Vf for any arbitrary part (with known or predicted thermal history) is spanned. 2(b) A database of these possible microstructures is generated, 2(c) the database include precomputed interaction tensors comprising the training stage of the reduced order mechanical model.

FIG. 91 shows for each macroscale material point (element, in this case), the thermal history and strain history are passed to a microscale solver; a microstructure is selected from the database based on the thermal history, and deformation boundary conditions are applied according to the strain history, according to embodiments of the invention. A crystal plasticity based microscale solution is computed, and a homogenized response (e.g., the l-norm of the fatigue indicating parameter, if a fatigue problem is chosen) is returned to the macroscale.

FIG. 92 shows (panel a) example component geometry specification (ASTM E606 fatigue specimen) (panel b) the two meshes and (panel c) details of the specimen meshes, including the difference between stress (left) and thermal (right) meshes, according to embodiments of the invention.

FIG. 93 shows the continuous radius fatigue test specimen, showing the thermal processing predictions partway through the build according to embodiments of the invention.

FIG. 94 shows estimated fatigue lives for multiple realizations of the fatigue coupon, run at different applied strain amplitudes, mimicking experimental conditions, according to embodiments of the invention. Two different processing conditions (conditions 1 and 2 in Table 6-2) were modeled.

FIG. 95 shows multiscale features of woven composites from microscale to macroscale according to embodiments of the invention.

FIG. 96 shows multiscale computational challenge of woven composite structures three scales; complexity, microstructure according to embodiments of the invention.

FIG. 97 shows SCA scheme for woven composites according to embodiments of the invention.

FIG. 98 shows concurrent multiscale simulation framework according to embodiments of the invention.

FIG. 99 shows the microstructure of plain weave composites according to embodiments of the invention.

FIG. 100 shows an UD RVE model according to embodiments of the invention

FIG. 101 shows clustering process and results of matrix with 256 clusters according to embodiments of the invention. Dimensions are given in FIG. 99.

FIG. 102 shows clustering process and results of yarns with 64 clusters according to embodiments of the invention. For each yarn, clustering is performed first based on local orientation. The resulting clusters are refined further using strain concentration Am tensor.

FIG. 103 shows the prediction results given by FEM and SCA (The SCA-64-128 indicates 64 clusters in matrix and 128 clusters in the yarns) according to embodiments of the invention

FIG. 104 shows the prediction results given by FEM and SCA according to embodiments of the invention.

FIG. 105 shows Hill yield surface calculation workflow according to embodiments of the invention. The 3D yield surfaces are plotted against three normal stress components and three shear stress components. For the plot against normal stress components, the cross section where sigma_zz=0 is illustrated. For the plot against shear stress components, the cross section where sigma_xy=0 is plotted.

FIG. 106 shows convergence study for different RVE sizes according to embodiments of the invention.

FIG. 107A shows geometry and mesh of the T-shaped hooking structure: geometry and boundary conditions, according to embodiments of the invention.

FIG. 107B shows geometry and mesh of the T-shaped hooking structure: mesh model and reduced order model, according to embodiments of the invention.

FIG. 108 shows Simulation results of the T-shaped hooking structure.

FIG. 109 shows (panel a) An idealized MVE of eight cubic grains with randomly assigned crystallographic orientation colored according to the inverse pole FIG. 109 in (b); the Euler angles for grain 1 to 8 are (15.2,65.0,184.0), (313.5,28.5,35.7), (102.1,171.5,245.7), (198.3,54,7,340.8), (288.2,1.1,53.1), (300.1,126.2,248.0), (325.7,133.5,39.6), and (47.2,127.1,196.3); (panel b) Inverse pole (IPF) color map showing the orientations in the MVE, with a z-face normal as the reference direction and three of the Miller indices in a hexagonal close packed lattice cell; (panel c) Elastic strain component E11 calculated using FEM with pure elasticity at 0.02 overall strain under uniaxial tension in the x-direction; (panel d) 128 clusters, 16 clusters/grain, obtained from (panel c) using the k-means clustering method, according to embodiments of the invention.

FIG. 110 shows (panel a) macroscale stress-strain curves calculated using CPFEM (in black), CPFFT (in red) with different mesh; (panel b) a closer look at the dashed rectangle of (a); (panel c) Macroscale stress-strain curves calculated using CPSCA (in blue) with increasing number of clusters per grain where the high-fidelity CPFEM and CPFFT results are shown as reference; and (panel d) a closer look at the dashed rectangle of (panel c), according to embodiments of the invention. The green shading indicates an area within 5% of the reference solution and the red within 1% of the reference solution.

FIG. 111 shows (panel a) The errors of 0.2% offset and 0.4% offset stress as a function of number clusters used per grain; (panel b) CPU time versus number of clusters (or voxels) per grain for CPSCA compared with using CPFEM and CPFFT, according to embodiments of the invention. Note that all simulations were run on Intel Haswell E5-2680v32.5 GHz processors.

FIG. 112 shows multiple realizations of the reconstruction-computation-evaluation loop generates a microstructure-property database according to embodiments of the invention.

FIG. 113 shows example MVEs with about 90 grains and corresponding grain size histograms according to embodiments of the invention. The grains are sampled with (panel a) no texture, (panel b) (0,0,0) preferred, (panel c) (90,0,0) preferred, and (panel d) (90,90,0) preferred, all of which are colored by the inverse pole of z-direction.

FIG. 114 shows predicted stress-strain curves of 50 MVEs each sampled with (panel a) no texture, (panel b) (0,0,0) preferred, (panel c) (90,45,0) preferred, and (panel d) (90,90,0) preferred, according to embodiments of the invention.

FIG. 115 shows distribution plots of predicted effective Young's modulus (panel a) and 0.2% offset effective yield strength (panel b) of different texture cases. Red lines show mean and one standard deviation spread, according to embodiments of the invention.

FIG. 116 shows example MVEs with increasing average grain size shows (panel a) ESD=13.2 μm, (panel b) ESD=19.7 μm, (panel c) ESD=26.6 μm, (panel d) ESD=35.9 μm, according to embodiments of the invention. The grains are colored by IPF.

FIG. 117 shows distribution plots of predicted effective Young's modulus (panel a) and 0.2% offset yield strength (panel b) with different averaged grain size, according to embodiments of the invention.

FIG. 118 shows a schematic of (panel a) the rolling process and (panel b) its concurrent multiscale simulation method, according to embodiments of the invention. Given deformation gradient F0 for each material point in the part, the first Piola-Kirchoff stress P0 is obtained by solving a polycrystalline MVE problem.

FIG. 119 shows contour of shear stress component σ12 (unit: MPa) at rolling time=0.08 seconds plotted on the deformed configuration, simulated with 3D elements (a and b) and plane strain elements (c), according to embodiments of the invention. 3D simulation predicts lower deformation in the rolling direction and higher extreme shear stress.

FIG. 120 shows history of σ12 for each integration point of the three elements indicated in panel (panel c) of FIG. 119, according to embodiments of the invention. The shear stress value of elements closer to the contact region tend to alternate more times and with higher amplitude.

FIG. 121 shows snapshots of macroscale shear stress contour, and microscale equivalent plastic strain and (0001) pole figures associated with the three elements, according to embodiments of the invention. Away from the contact surface the associated MVE experiences less rotation and shear, and more compression and plastic strain.

FIG. 122 shows uniaxially tensile (panel a) and compressive (panel b) stress-strain curves for Epon 825 deformed at a strain rate of 5×108 s−1 for different temperatures, according to embodiments of the invention.

FIG. 123 shows yield surfaces obtained for different temperatures according to embodiments of the invention, where the points are simulation data and the lines are theoretical prediction using Eq. (9-1).

FIG. 124 shows effect of crosslink degree and component ratio on the stress-strain behavior by using epoxy 3501-6 as a model system, according to embodiments of the invention.

FIG. 125 shows interphase property characterization, according to embodiments of the invention. (panel a) Schematic of the cross-section including the interphase region (yellow). (panel b) Variation of Young's modulus or strength inside the interphase region.

FIG. 126 shows Prepreg's deformation mechanisms during preforming, according to embodiments of the invention.

FIG. 127 shows the experiment setting for the uniaxial tension and bias-extension tests, according to embodiments of the invention.

FIG. 128 shows engineering strain-stress curves from the uniaxial tension tests, according to embodiments of the invention.

FIG. 129 shows bias-extension test results for different specimen sizes: (panel a) original load-displacement curves and (panel b) normalized load-displacement curves, according to embodiments of the invention.

FIG. 130 shows bias-extension test results for (panel a) different temperatures and (panel b) different tensile rates, according to embodiments of the invention.

FIG. 131 shows validation for the kinematic assumption of the bias-extension tests from (panel a) Green strain field obtained from DIC and (panel b) average Green strain comparison in the central region.

FIG. 132 shows (panel a) schematic of the bending test setup and (panel b) the shape of prepreg at 50° C., according to embodiments of the invention.

FIG. 133 shows experimental setup for the prepreg-prepreg interaction test apparatus, according to embodiments of the invention.

FIG. 134 shows schematic of the interaction measuring experimental apparatus, according to embodiments of the invention.

FIG. 135 shows force and interaction factor results from the test under the conditions of 70° C., 5 mm/s, and 0/90/0/90 fiber orientation, according to embodiments of the invention.

FIG. 136 shows steady-state interaction factor in a periodical variation subjected to the test conditions of 50° C., 15 mm/s, for the 0/90/0/90 fiber orientation, according to embodiments of the invention.

FIG. 137 shows interaction and stick-slip strength at various temperatures subjected to different (panel a) relative motion speeds and (panel b) fiber orientations, according to embodiments of the invention.

FIG. 138 is an illustration of the prepreg structure via (panel a) real product photo and (panel b) model generated by the software TexGen, according to embodiments of the invention.

FIG. 139 shows seometry and forces of the simulated two 2×2 twill fabric interface, according to embodiments of the invention.

FIG. 140 shows (panel a) experimental and numerical interaction factor comparison at various temperatures and 10 mm/s; and (panel b) a zoom-in to 60° c. and 70° c. for clear illustration, according to embodiments of the invention.

FIG. 141 shows experimental and numerical interaction factor comparison at various speeds and 60° C. temperature, according to embodiments of the invention. The points are moved away with the input speeds artificially to better differentiate between the data.

FIG. 142 shows fast Fourier transformation (FFT) results of the numerical and experimental data, according to embodiments of the invention.

FIG. 143 shows stress analysis in the non-orthogonal material model, according to embodiments of the invention.

FIG. 144 shows calculation flowchart of the LS-DYNA MAT 293, according to embodiments of the invention.

FIG. 145 shows TexGen rough geometry model with the thickness of the prepreg as 1.2 mm, according to embodiments of the invention: (panel a) the structure, and (panel b) the cross-section of the corresponding mesh.

FIG. 146 shows prepreg RVE compression in FE software Abaqus: two rigid plates are introduced to adjust the RVE thickness, according to embodiments of the invention.

FIG. 147 is an illustration of the yarn cross-section deformation upon compression along the width direction: (panel a) real material deformation mode, (panel b) FE deformation mode with transversely isotropic material model, and (panel c) FE deformation mode with decoupled material model in FE, according to embodiments of the invention.

FIG. 148 shows one bias-extension RVE simulation example: (panel a) illustration of the von Mises stress contour on the RVE; (panel b) comparison of the simulation and experimental true shear stress, according to embodiments of the invention.

FIG. 149 shows modular Bayesian calibration, according to embodiments of the invention: The approach has four stages and enables estimating the potential simulator bias as well as the joint posterior distribution of the calibration parameters.

FIG. 150 shows marginal posterior distributions on the calibration parameters, according to embodiments of the invention: The posterior and prior are indicated with solid blue and dotted red lines, respectively.

FIG. 151 shows posterior mean of the responses, according to embodiments of the invention: (panel a) normal stress as a function of normal true strain along the yarns for two different shear angles; (panel b) shear stress as a function of normal true strain along the yarns for two different fabric shear angles; (panel c) uniaxial tension test used in calibration vs. our predictions; and (panel d) bias extension test which is not used in calibration vs. our predictions.

FIG. 152 shows flowchart of the developed multiscale preforming simulation method, according to embodiments of the invention: The Bayesian calibration utilizes the RVE and experiments to obtain the yarn properties and the mesoscale stress emulator. The stress emulator is then implemented into the non-orthogonal material model for macroscopic preformation simulation.

FIG. 153 shows schematic of different scales in CFRP, according to embodiments of the invention.

FIG. 154 shows microscopy image of unidirectional UD cross-section, according to embodiments of the invention.

FIG. 155 shows UD RVE isometric view with fiber in black and matrix in grey, according to embodiments of the invention.

FIG. 156 shows woven RVE geometry, according to embodiments of the invention.

FIG. 157 shows an illustration of yarn angle, yarn fiber volume fraction, and fiber misalignment in yarn, according to embodiments of the invention.

FIG. 158 shows effect of yarn angle on components of woven stiffness matrix, according to embodiments of the invention.

FIG. 159 shows Illustration of clusters in the fiber phase and the matrix phase, according to embodiments of the invention.

FIG. 160 shows schematic of UD concurrent multiscale modeling, according to embodiments of the invention.

FIG. 161 shows (panel a) UD coupon modeling geometry and (panel b) boundary condition, according to embodiments of the invention.

FIG. 162 shows Von Mises stress contour of the coupon model and local RVEs a) before crack initiation b) after crack formation, according to embodiments of the invention.

FIG. 163 shows (panel a) UD hat-section crash model setup; (panel b) impactor force vs. time plot, according to embodiments of the invention.

FIG. 164 shows schematic of concurrent simulation of UD dynamic 3-pt bending, according to embodiments of the invention.

FIG. 165 shows 3-scale concurrent multiscale modeling setup, according to embodiments of the invention.

FIG. 166 shows geometry of woven RVE in concurrent multiscale modeling, according to embodiments of the invention.

FIG. 167 shows engineering shear stress vs. engineering shear strain of 3-scale and 2-scale single element simple shear, according to embodiments of the invention.

FIG. 168 shows woven bias tension test setup, according to embodiments of the invention.

FIG. 169 shows (panel a) woven bias sample σ22 contour with 0.4 mm applied displacement to the top tab; (panel b) σ22 vs ε22 curves of concurrent biaxial tension and three sets of test data, according to embodiments of the invention.

FIG. 170 shows multiscale structure schematic view of a four-scale woven fiber composite with polymer matrix, according to embodiments of the invention. In computational modeling of this structure, each integration point at any scale is a realization of a structure at the next finer scale.

FIG. 171 shows demonstration of our approach for s two-scale structure, according to embodiments of the invention: spatial random processes (SRPs) are employed for generating spatial variations that are connected through the top-down sampling procedure.

FIG. 172 shows the macroscopic cured woven laminate structure studied in our work, according to embodiments of the invention. (panel a) the deformed structure, where the light blue lines indicate the fiber orientation, and the dimensions are scaled for a clearer representation; (panel b) the deterministic spatial variations of yarn angle obtained from simulating a perfectly manufactured composite; (panel c) Von Mises stress field corresponding to Case 9; (panel d) the random spatial variations of yarn angle corresponding to one of the realizations of Case 1; and (panel e) the random spatial variations of θ1 corresponding to one of the realizations of Case 3.

FIG. 173 shows fiber misalignment angles, according to embodiments of the invention. The zenith and azimuth angles characterize the fiber misalignment angle with respect to the local orthogonal frame on the yarn cross-section.

FIG. 174 shows coupling the uncertainty sources across the scales, according to embodiments of the invention: The spatial variations of v and θ at the macroscale are connected to those at the finer scales. For brevity, the coupling is illustrated only for the average values for the two quantities (i.e., the mean of the RFs: β=[βv, βθ]=[v2, θ2]).

FIG. 175 shows effect of average values on the effective moduli of a woven RVE, according to embodiments of the invention: (panel a) Effect of fiber volume fraction and, (panel b) Effect of misalignment. Each point on these figures indicates the average value over 20 simulations. The Case IDs in (panel b) are defined in Table 9-17. The reference solution refers to a case where there is no misalignment.

FIG. 176 shows prediction error as a function of the number of training samples, according to embodiments of the invention: As the number of training samples increases, the accuracy of the MRGP metamodel in predicting the elements of the stiffness matrix of the mesoscale woven RVE increases.

FIG. 177 shows screenshots of graphical user interfaces, according to embodiments of the invention: (panel a) optimal Latin hypercube sampling user interface. (panel b) Gaussian process modeling user interface.

FIG. 178 shows global and local response of the macrostructure to the spatial variations, according to embodiments of the invention: (panel a) reaction force, (panel b) mean stress at the mid-section, and (panel c) standard deviation of the stress field at the mid-section. for cases 1 through 8 in panels (b) and (c), the curves are based on 20 independent simulations.

FIG. 179 shows (panel a) a typical fiber volume fraction (VF) contour, yellow color represents high VF (about 0.6), and blue color represents low VF (about 0.4); and (panel b) the use of the location of splitting lines to represent the non-stationarity, according to embodiments of the invention.

FIG. 180 shows (panel a) probability density estimate of the inter-line distances of 3 of the samples; and (panel b) Probability density estimates of 3 batches of the samples (each batch contains 10 samples), according to embodiments of the invention.

FIG. 181 shows two examples of reconstructions, according to embodiments of the invention.

FIG. 182 shows the segmented linear regression process, according to embodiments of the invention.

FIG. 183 shows an illustration of the local waviness found by the proposed method, according to embodiments of the invention.

FIG. 184 shows fiber waviness characterization of a large image, according to embodiments of the invention.

FIG. 185 shows some reconstructions and their corresponding periodograms, according to embodiments of the invention.

FIG. 186 shows geometry of the double-dome punch and the binder, according to embodiments of the invention.

FIG. 187 shows double-dome preforming test setup, according to embodiments of the invention: (panel a) the press for the preforming, and (panel b) the prepreg temperature history plot. The plot indicates that the prepreg temperature drops rapidly from the initial 70° C. to around 23° C. when it is placed under the press.

FIG. 188 shows double-dome preforming simulation setup, according to embodiments of the invention.

FIG. 189 shows simulation and experimental results comparison of deformed geometry and yarn angle distribution for double-dome preformed part of ±45° single layer woven prepreg.

FIG. 190 shows preforming simulation, according to embodiments of the invention: (panel a) final part shapes and yarn angle distributions, and (panel b) punch force comparison. In panel (a), A-E points indicate yarn angle measuring positions.

FIG. 191 shows double-dome comparison with different initial prepreg layer orientations, according to embodiments of the invention: (panel a) 0/90, (panel b) −45/+45, and (panel c) 0/90/−45/+45. The simulation results are shown in the top half while the experimental ones are in the bottom half. The silver lines on the experimental results indicate the directions of the warp and weft yarns.

FIG. 192 shows double-dome warp-weft yarn angle distribution results, according to embodiments of the invention: (panel a) sampling points on the 0/90 prepreg layer, (panel b) sampling points on the −45/+45 sampling points, (panel c) angle comparison for the single layer 0/90 fiber orientation preforming, (panel d) angle comparison for the single layer −45/+45 fiber orientation preforming, (panel e) angle comparison for the 0/90 fiber orientation layer in the 0/90/−45/+45 double layer preforming, and (panel f) angle comparison for the −45/+45 fiber orientation layer in the 0/90/−45/+45 double layer preforming.

FIG. 193 shows folding of the prepreg after low temperature preforming especially at the edge, according to embodiments of the invention.

DETAILED DESCRIPTION OF THE INVENTION

The present invention will now be described more fully hereinafter with reference to the accompanying drawings, in which exemplary embodiments of the present invention are shown. The present invention may, however, be embodied in many different forms and should not be construed as limited to the embodiments set forth herein. Rather, these embodiments are provided so that this disclosure will be thorough and complete, and will fully convey the scope of the invention to those skilled in the art. Like reference numerals refer to like elements throughout.

The terms used in this specification generally have their ordinary meanings in the art, within the context of the invention, and in the specific context where each term is used. Certain terms that are used to describe the invention are discussed below, or elsewhere in the specification, to provide additional guidance to the practitioner regarding the description of the invention. For convenience, certain terms may be highlighted, for example using italics and/or quotation marks. The use of highlighting and/or capital letters has no influence on the scope and meaning of a term; the scope and meaning of a term are the same, in the same context, whether or not it is highlighted and/or in capital letters. It will be appreciated that the same thing can be said in more than one way. Consequently, alternative language and synonyms may be used for any one or more of the terms discussed herein, nor is any special significance to be placed upon whether or not a term is elaborated or discussed herein. Synonyms for certain terms are provided. A recital of one or more synonyms does not exclude the use of other synonyms. The use of examples anywhere in this specification, including examples of any terms discussed herein, is illustrative only and in no way limits the scope and meaning of the invention or of any exemplified term. Likewise, the invention is not limited to various embodiments given in this specification.

It will be understood that, although the terms first, second, third, etc. may be used herein to describe various elements, components, regions, layers and/or sections, these elements, components, regions, layers and/or sections should not be limited by these terms. These terms are only used to distinguish one element, component, region, layer or section from another element, component, region, layer or section. Thus, a first element, component, region, layer or section discussed below can be termed a second element, component, region, layer or section without departing from the teachings of the present invention.

It will be understood that, as used in the description herein and throughout the claims that follow, the meaning of “a”, “an”, and “the” includes plural reference unless the context clearly dictates otherwise. Also, it will be understood that when an element is referred to as being “on,” “attached” to, “connected” to, “coupled” with, “contacting,” etc., another element, it can be directly on, attached to, connected to, coupled with or contacting the other element or intervening elements may also be present. In contrast, when an element is referred to as being, for example, “directly on,” “directly attached” to, “directly connected” to, “directly coupled” with or “directly contacting” another element, there are no intervening elements present. It will also be appreciated by those of skill in the art that references to a structure or feature that is disposed “adjacent” to another feature may have portions that overlap or underlie the adjacent feature.

It will be further understood that the terms “comprises” and/or “comprising,” or “includes” and/or “including” or “has” and/or “having” when used in this specification specify the presence of stated features, regions, integers, steps, operations, elements, and/or components, but do not preclude the presence or addition of one or more other features, regions, integers, steps, operations, elements, components, and/or groups thereof.

Furthermore, relative terms, such as “lower” or “bottom” and “upper” or “top,” may be used herein to describe one element's relationship to another element as illustrated in the figures. It will be understood that relative terms are intended to encompass different orientations of the device in addition to the orientation shown in the figures. For example, if the device in one of the figures is turned over, elements described as being on the “lower” side of other elements would then be oriented on the “upper” sides of the other elements. The exemplary term “lower” can, therefore, encompass both an orientation of lower and upper, depending on the particular orientation of the figure. Similarly, if the device in one of the figures is turned over, elements described as “below” or “beneath” other elements would then be oriented “above” the other elements. The exemplary terms “below” or “beneath” can, therefore, encompass both an orientation of above and below.

Unless otherwise defined, all terms (including technical and scientific terms) used herein have the same meaning as commonly understood by one of ordinary skill in the art to which the present invention belongs. It will be further understood that terms, such as those defined in commonly used dictionaries, should be interpreted as having a meaning that is consistent with their meaning in the context of the relevant art and the present disclosure, and will not be interpreted in an idealized or overly formal sense unless expressly so defined herein.

As used in this disclosure, “around”, “about”, “approximately” or “substantially” shall generally mean within 20 percent, preferably within 10 percent, and more preferably within 5 percent of a given value or range. Numerical quantities given herein are approximate, meaning that the term “around”, “about”, “approximately” or “substantially” can be inferred if not expressly stated.

As used in this disclosure, the phrase “at least one of A, B, and C” should be construed to mean a logical (A or B or C), using a non-exclusive logical OR. As used herein, the term “and/or” includes any and all combinations of one or more of the associated listed items.

The methods and systems will be described in the following detailed description and illustrated in the accompanying drawings by various blocks, components, circuits, processes, algorithms, etc. (collectively referred as “members”). These members may be implemented using electronic hardware, computer software, or any combination thereof. Whether such elements are implemented as hardware or software depends upon the particular application and design constraints imposed on the overall system. By way of example, a member, or any portion of an member, or any combination of members may be implemented as a “processing system” that includes one or more processors. Examples of processors include microprocessors, microcontrollers, graphics processing units (GPUs), central processing units (CPUs), application processors, digital signal processors (DSPs), reduced instruction set computing (RISC) processors, systems on a chip (SoC), baseband processors, field programmable gate arrays (FPGAs), programmable logic devices (PLDs), state machines, gated logic, discrete hardware circuits, and other suitable hardware configured to perform the various functionality described throughout this disclosure. One or more processors in the processing system may execute software. Software shall be construed broadly to mean instructions, instruction sets, code, code segments, program code, programs, subprograms, software components, applications, software applications, software packages, routines, subroutines, objects, executables, threads of execution, procedures, functions, etc., whether referred to as software, firmware, middleware, microcode, hardware description language, or otherwise.

Accordingly, in one or more example embodiments, the functions described may be implemented in hardware, software, or any combination thereof. If implemented in software, the functions may be stored on or encoded as one or more instructions or code on a computer-readable medium. Computer-readable media includes computer storage media. Storage media may be any available media that can be accessed by a computer. By way of example, and not limitation, such computer-readable media can comprise a random-access memory (RAM), a read-only memory, an electrically erasable programmable read-only memory (EEPROM), optical disk storage, magnetic disk storage, other magnetic storage devices, combinations of the aforementioned types of computer-readable media, or any other medium that can be used to store computer executable code in the form of instructions or data structures that can be accessed by a computer.

The description below is merely illustrative in nature and is in no way intended to limit the invention, its application, or uses. The broad teachings of the invention can be implemented in a variety of forms. Therefore, while this invention includes particular examples, the true scope of the invention should not be so limited since other modifications will become apparent upon a study of the drawings, the specification, and the following claims. For purposes of clarity, the same reference numbers will be used in the drawings to identify similar elements. It should be understood that one or more steps within a method may be executed in different order (or concurrently) without altering the principles of the invention.

Microstructured materials are aggregates of individual components. In multi-component materials, the components, which are considered as material building blocks, aggregate or self-assemble to form a complex structures or conformations at multiple scales. Multiscale modeling methods attempt to faithfully capture the emergent complex behaviors on several length- and time-scales.

One of the objectives of this invention is to provide data-driven representation and clustering discretization methods and systems, which is a data-driven domain decomposition approach that is suitable to accelerate the numerical simulation of the microstructure responses. The concept here is similar to the Lebesgue Integral, illustrated in FIG. 2, which depends on grouping nearby response (y-axis) rather than nearby domain (x-axis) used for the Riemann Integral. This decomposition approach significantly reduces overall computational complexity of a problem, especially in multiscale analysis.

In one aspect of the invention, a method for design optimization and/or performance prediction of a material system includes generating a representation of the material system at a number of scales, wherein the representation at a scale comprises microstructure volume elements (MVE) that are building blocks of the material system at said scale; collecting data of response fields of the MVE computed from a material model of the material system over a predefined set of material properties and boundary conditions; applying machine learning to the collected data of response fields to generate clusters that minimize a distance between points in a nominal response space within each cluster; computing an interaction tensor of interactions of each cluster with each of the other clusters; and solving an governing partial differential equation (PDE) using the generated clusters and the computed interactions to result in a response prediction that is usable for the design optimization and/or performance prediction of the material system.

In one embodiment, the method further comprises passing the resulting response prediction to a next coarser scale as an overall response of that building block, and iterating the process until a final scale is reached.

In one embodiment, the building blocks are defined by material properties and structural descriptors obtained by modeling or experimental observations and encoded in a domain decomposition of structures for identifying locations and properties of each phase within the building blocks.

In one embodiment, the structural descriptors comprise characteristic length and geometry.

In one embodiment, the boundary conditions are chosen to satisfy the Hill-Mandel condition.

In one embodiment, the collected data of response fields comprise a strain concentration tensor, a deformation concentration tensor, stress tensor (e.g., PK-I stress, Cauchy stress), plastic strain tensor, thermal gradient, or the like.

In one embodiment, the machine learning comprises unsupervised machine learning and/or supervised machine learning.

In one embodiment, the machine learning is performed with a self-organizing mapping (SOM) method, a k-means clustering method, or the like.

In one embodiment, the clusters are generated by marking all material points that have the same response field within the representation of the material system with a unique ID and grouping material points with the same ID.

In one embodiment, the generated clusters are a reduced representation of the material system, which reduces the number of degrees of freedom required to represent the material system.

In one embodiment, the generated clusters are a reduced order MVE of the material system.

In one embodiment, the computed interaction tensor is for all pairs of the clusters.

In one embodiment, said computing of the interaction tensor is performed with fast Fourier transform (FFT), numerical integration, or finite element method (FEM).

In one embodiment, the PDE is a Lippmann-Schwinger equation. In one embodiment, said solving of the PDE is performed with arbitrary boundary conditions and/or material properties.

In one embodiment, the collected data of response fields, the generated clusters, and/or the computed interaction tensor are saved in one or more material system databases.

In one embodiment, said solving of the PDE is performed in real-time by accessing the one or more material system databases for the generated clusters and the computed interactions.

In another aspect of the invention, a method for design optimization and/or performance prediction of a material system includes performing an offline data compression, wherein original microstructure volume elements (MVE) of building blocks of the material system are compressed into clusters, and an interaction tensor of interactions of each cluster with each of the other clusters is computed; and solving an governing PDE using the clusters and the computed interactions to result in a response prediction that is usable for the design optimization and/or performance prediction of the material system.

In one embodiment, the method further includes passing the resulting response prediction to a next coarser scale as an overall response of that building block, and iterating the process until a final scale is reached.

In one embodiment, the building blocks are defined by material properties and structural descriptors obtained by modeling or experimental observations and encoded in a domain decomposition of structures for identifying locations and properties of each phase within the building blocks.

In one embodiment, the structural descriptors comprise characteristic length and geometry.

In one embodiment, when more than one scale is involved with the reduced order response prediction, the method is named Multiresolution Clustering Analysis (MCA).

In one embodiment, the boundary conditions are chosen to satisfy the Hill-Mandel condition.

In one embodiment, said performing the offline data compression comprises collecting data of response fields of the MVE computed from a material model of the material system over a predefined set of material properties and boundary conditions; applying machine learning to the collected data of response fields to generate clusters that minimize a distance between points in a nominal response space within each cluster; and computing the interaction tensor is for all pairs of the clusters.

In one embodiment, the collected data of response fields comprise a strain concentration tensor, a deformation concentration tensor, stress tensor (e.g., PK-I stress, Cauchy stress), plastic strain tensor, thermal gradient, or the like.

In one embodiment, the machine learning comprises unsupervised machine learning and/or supervised machine learning.

In one embodiment, the machine learning is performed with an SOM method, a k-means clustering method, or the like.

In one embodiment, the clusters are generated by marking all material points having the same response field within the representation of the material system with a unique ID and grouping material points with the same ID.

In one embodiment, the clusters are a reduced representation of the material system, which reduces the number of degrees of freedom required to represent the material system.

In one embodiment, the clusters are a reduced order MVE of the material system.

In one embodiment, said computing the interaction tensor is performed with FFT, numerical integration, or FEM.

In one embodiment, the PDE is a Lippmann-Schwinger equation. h one embodiment, said solving the PDE is performed with arbitrary boundary conditions and material properties.

In one embodiment, the collected data of response fields, the generated clusters, and/or the computed interaction tensor are saved in one or more material system databases.

In one embodiment, said solving the PDE is performed with online accessing the one or more material system databases for the generated clusters and the computed interactions.

In yet another aspect, the invention relates to a material system database usable for conducting efficient and accurate multiscale modeling of a material system, In one embodiment, the material system database includes clusters for a plurality of material systems, each of which groups all material points having a same response field within MVE of a respective material system with a unique ID; interaction tensors, each of which represents interactions of all pairs of the clusters for the respective material system; and response predictions computed based on the clusters and the interaction tensors.

In one embodiment, the clusters are generated by applying machine learning to data of response fields of the MVE computed from a material model of the respective material system over a predefined set of material properties and boundary conditions.

In one embodiment, the interaction tensors are computed with FFT, numerical integration, or FEM.

In one embodiment, the responses predictions are obtained by solving a governing PDE using the clusters and the computed interactions. In one embodiment, the responses predictions comprise at least effective stiffness, yield strength, thermal conductivity, damage initiation, and FIP.

In one embodiment, the material system database is configured such that some of the response predictions are assigned as a training set for training a different machine learning that connects processes/structures to responses/properties of the material system directly without going through the clustering and interaction computing processes at all; and some or all of the remaining response predictions are assigned as a validation set for validating the efficiency and accuracy of the multiscale modeling of the material system.

In one embodiment, the material system database is generated with predictive reduced order models. In one embodiment, the predictive reduced order models comprise a self-consistent clustering analysis (SCA) model, a virtual clustering analysis (VCA) model, and/or an FEM clustering analysis (FCA) model.

In one embodiment, the material system database is updatable, editable, accessible, and searchable.

In a further aspect, the invention relates to a method of applying the above-disclosed material system database for design optimization and/or performance prediction of a material system. In one embodiment, the method includes training a neural network with data of the material system database; and predicting real-time responses during a loading process performed using the trained neueral network, wherein the real-time responses are used for the design optimization and/or performance prediction of a material system.

In one embodiment, the method further includes performing a topology optimization to design a structure with microstructure information.

In one embodiment, the neural network comprises a feed forward neural network (FFNN) and/or a convolutional neural network (CNN).

In one aspect, the invention relates to a non-transitory tangible computer-readable medium storing instructions which, when executed by one or more processors, cause a system to perform the above-disclosed methods for design optimization and/or performance prediction of a material system.

In another aspect, the invention relates to a computational system for design optimization and/or performance prediction of a material system. In one embodiment, the computational system includes one or more computing devices comprising one or more processors; and a non-transitory tangible computer-readable medium storing instructions which, when executed by the one or more processors, cause the one or more computing devices to perform the above-disclosed method for design optimization and/or performance prediction of a material system.

The advantages and specific applications of the invented methods and systems are briefed as follows, while the details of them are discussed in EXAMPLES 1-9 following the section.

A methodology for constructing microstructure material databases for fast microstructure-derived response prediction has been developed. The method takes materials data and compresses it using unsupervised machine learning to create a “clustering discretization” in effect a specially designed database that is suitable to conduct efficient and accurate multiscale modeling of arbitrary material systems. Because the method described here relies on capturing and combining the fundamental building blocks of the material and its response, any complex and/or hierarchical material systems can be accurately and efficiently modeled, thus, the method is material agnostic. The method described here can also be applicable to prediction of many physical phenomena that share similar underlying mathematical descriptions. While the examples provided are related to the material behavior of solids subjected to mechanical loads, this should not be thought of as a fundamental restriction; the method could well be applied to predict a range of effective physics, including, but not limited to, electric and magnetic properties. In effect, the method might be used to solve arbitrary computational homogenization problems, or problems involving the solution of partial differential equations.

FIG. 1 illustrates schematically an overall flow of data through the system according to embodiments of the invention, where heavy bordered boxes, e.g., 110, 120, 130, 140, 150, are operations, typically including computer codes, and these operations act upon various data contained within the parallelograms, e.g., 115, 125, 135, 145, 155, and typically stored on disk or in memory depending upon how the operations are embodied. In the more complex operations, information or steps commonly used to produce a desired operation are listed; these appear as numbered, round-cornered boxes. The system starts with a representation of the material system at a finite number of scales. Any particular scale is composed of fundamental building blocks, the size of which is defined by a characteristic length. The composition of the building blocks is defined by modeling or experimental observations. These are then encoded in a detailed spatial decomposition of the structure (sometimes called a “mesh”), used to describe the location and properties of each phase within the building blocks. Generation of this high-resolution description is termed as microstructure generation and may be relevant to any scale. A predefined set (usually carefully selected and simplified) of material properties and boundary conditions are supplied to a direct numerical solver to compute nominal response fields. Unsupervised machine learning is applied to these fields to generate clusters that minimize the distance between points in nominal response space within each cluster. This produces a reduced representation of the material building block. The interaction between each cluster (or, the influence that a unit load applied to one cluster has on other clusters) can be computed and stored. Using precomputed clusters and interactions, the solution to the relevant governing partial differential equation (PDE) with arbitrary boundary conditions and material properties (not necessarily the simplified ones used to compute clusters) is fast in the response prediction stage. The results of the response prediction are passed to the next coarser scale, as the overall response of that building block, and the process thus proceeds.

A multiscale analysis is envisioned to include models that capture material behavior at each relevant length scale passing information to the next higher length scale—until the final scale is reached. This appears on the far left of FIG. 1

The next step shown in FIG. 1 involves generation of an MVE, or material representation. A reasonable MVE, or set of MVEs depending on the number of scales involved, for any given problem might be derived from many different sources. This method can be used with MVEs derived from process simulations EXAMPLE 2, purely conceptual or synthetically generated MVEs from statistics EXAMPLES 2-3 and 11, and with MVEs taken directly from experimental images in a variety of ways, see, e.g., EXAMPLES 5-6. An example of MVE construction and processing for metals is given in FIG. 8, where a combination of statistical reconstruction and direct imaging is used to create an MVE.

For an example in composite materials, in the simplest form, the final scale might be a component (scale 0, the largest) that is built with unidirectional carbon fiber reinforced composite (scale 1). At scale 1, a microstructure model to compute physical material responses is required. Once computed, this response is provided to scale 0 as the element-level behavior. This is illustrated in FIG. 5. To capture its response, the microstructure is modeled as a high-fidelity MVE. Each MVE contains millions of voxel elements, as shown in FIG. 4. One might envision how this same scheme might be extended to scale 2, e.g., for a woven composite including unidirectional fiber in each yarn, as shown in FIG. 7. For arbitrary material systems, this scale decomposition might continue to scale N.

Metals are another example of materials displaying functional and structural hierarchy. In one simple case, metal might be represented by a homogenized behavior at scale 0, while scale 1 describes the behavior at the granular level. The collective behavior of many grains works to generate the homogenized response observed in scale 0, see, e.g., EXAMPLE 2. Another might be that of defects (inclusions, voids) within each grain, and yet another the scale of precipitates and dislocations.

According to the invention, the core of the method is composed of the follow three steps, outlined as the top row in FIG. 1:

(1) Data collection using the high-fidelity MVE. This operation uses a high-fidelity approach with specifically crafted boundary conditions and properties. This gives an indicator of the material response of the MVE, for each MVE. The data collected for each MVE are called “response fields,” implying that there is not necessarily a unique choice of material response required. Often elastic response is desirable for its simplicity, but the method can use any response field deemed appropriate (for example, plastic strain is another possibility). This data is saved in a database.

(2) Perform unsupervised learning on the response data obtained in step (1). This process marks all points within the MVE with a unique ID, such as 1, 2, 3, etc. Thence the description of the MVE behavior can rely on the groups of points with the same ID, rather than the points themselves. We call each group a “cluster.” This is illustrated in FIG. 2. The mapping between Cluster ID and spatial point is stored in a database called “clusters.” The clusters must be space-filling, though each cluster need not be contiguous.

(3) Compute the interaction tensor for all pairs of clusters (rather than points). This process computes pairwise interaction tensors for all clusters. This allows one to compute behavior of each cluster when some forcing term is applied on the MVE. At this step, the original MVE has been completely replaced by the “compressed” MVE mathematically described with only clusters, not spatial points/elements.

Above three-step approach is depicted in further detail in FIG. 6. After this has been done, the data is read in by the prediction operator in the lower portion shown in FIG. 1. This approach leads to a drastically reduced computational cost for solving the MVE responses, as shown in Table 1, but the accuracy is preserved, as shown by an example in FIG. 9. The computational time comparison for cases given in FIG. 9 is provided in Table 1. With such a drastic reduction in computational time, practical modeling challenges requiring description of a material at multiple scales are overcome using desktop-workstation-level computational resources.

The reduction of complexity is illustrated in FIG. 73. The selection of the MVE size can be determined through a statistical analysis process. The interaction tensor computation can be accelerated using various formulations. As shown in Table 1, one can see that the model without any data compression is big and the simulation is costly to perform. The invented method significantly reduces the overall system complexity (e.g., as measured by degrees of freedom in the solution) and the computational cost.

TABLE 1 Comparison of computational time and average percent difference between the DNS and reduced order MVE calculation. Difference MVE to DNS MVE Type Complexity Computational Time Solution High fidelity MVE 36,000,000 200 hr. with 80 CPU cores 0% (DNS) voxels (57.6 M CPU-seconds) Reduced order 10 clusters Offline data collection, 8 clusters: MVE 8,280 s 3.83% Online prediction, 2 s with 1 16 clusters: CPU 1.46%

The superior efficiency can also be used to build a complex material microstructure response database using the reduced order model (ROM), e.g., EXAMPLE 1. One possible use for such a database is for training a physics based neural network, which provides almost instantaneous microstructure responses under arbitrary external loading condition, as shown schematically in FIG. 10. The physics based neural network can be used for structure optimization and design, as suggested in EXAMPLE 1. Moreover, the methodology described is suitable for materials and systems design for desired performance indices, as illustrated in FIG. 11.

FIG. 3 shows example material systems with two or more relevant length scales that could be modeled by the method and system.

FIG. 4 shows one exemplary MVE model of a unidirectional carbon fiber reinforced composite containing 36 million voxels. On the left, the blue cylinders represent carbon fiber, and the red is a polymetric epoxy resin matrix materials. The overall (homogenized) response to an applied load in the y-direction is plotted in the top right, and the local stress contours are plotted on the bottom right. The DNS process to compute this single load history takes about 200 hours using 80 computer processors.

FIG. 5 is an illustration of the 2-scale multiscale problem and its complexity using the traditional finite element method. At the macroscale, a mesh of a generic part subject to some boundary condition is shown. At each material point within these elements, a microscale response corresponding to the behavior of a generic subscale, shown in this example with spherical inclusions, is computed. Bottom: Illustration of two-scale multiscale clustering method and its capability in order reduction. In this illustration, each color represents order 1 degrees of freedom, rather than each blue-bounded region within the top illustration.

FIG. 6 shows an exemplary three-step “training” process, as one possible way to compute the clusters and interaction tensor for macroscale geometry (top) and microscale geometry (bottom).

FIG. 7 shows an exemplary three-scale FRP modeling framework. The macroscale model is a woven laminate composite model, built as a Finite Element mesh. Each integration point in the macroscale model is represented by a woven MVE, which can have different tow size, tow spacing, and tow angle (the one shown is 90° tow angle). Each integration point in the mesoscale model is represented by a UD MVE, which can have different fiber orientation, fiber volume fraction, and matrix-fiber interfacial strength. If all three scales are discretized with FE meshes, the total DOF for this full-field system could be 3.3×1015. Using the present method and replacing the mesoscale and the microscale with ROMs, the total DOF is reduced down to 1.6×106, a reduction of nine order of magnitude.

FIG. 8 shows that in certain embodiment, MVEs might be generated from a number of sources; this example for metallic materials shows grains measured using x-ray diffraction, reconstructed from a statistical description, and predicted from a processing model. Defects, voids in this case, can be from measurements or predictions, here two experimental methods (x-ray tomography and FIB-SEM serial sectioning) are given as examples. These input data are paired (spatially), and used for the MVE, resulting in response predictions based directly on experimental images.

In FIG. 9, left panel is the original DNS description of a UD MVE, center panel is cluster-based description (showing 2 clusters in the fiber phase, 8 in the matrix), visualized on the underlying voxel mesh, and right panel is numerical homogenization results for overall stress/strain. The DNS reference solution is shown in red with pointwise 5% error bars. Cf. the reduced order solutions with 8 clusters (blue) and 16 clusters (green).

Referring to FIG. 10, a physics guided NN may be “layered” on top of the MVE ROM: the NN is trained on a large database of rapidly-computed behavior. Once trained, this NN is thought to contain microstructural information similar to the ROM, but is much faster to evaluate. This alternative makes it practical to conduct microstructure-based structural optimization and design.

FIG. 11 shows an application example: a composite design. The composite structure can be designed with different microscale structure (e.g., fiber shape) and mesoscale structure (e.g., fiber orientation in each ply, fiber orientation and fiber shape in each ply, as well as different weave pattern for woven composite). Key performance indices (e.g., strength and maximum deformation under external load) for the composite structure can be predicted. If those indices do not meet the desired criterion, optimization routine will be called to update microscale and mesoscale structure in order to improve performance indices.

According to the invention, the process is fast, see, e.g., EXAMPLES 5-8, among others, as shown in Table 1.

It is also efficient, based on the following justification (EXAMPLES 5-8). The efficiency comes from the fact that the complexity of the problem is reduced significantly after the data compression. The results present in EXAMPLE 5, illustrate the saving of computational cost, as shown in Table 5-4. The present method provides considerable reduction in terms of computer memory needed and computational time needed, compared to the traditional approaches, such as FEM and FFT.

Further, the methos is also accurate, as demonstrated in Table 1 and shown in FIG. 9.

Moreover, the method can be applied to materials with one or more length scales, as provided in EXAMPLES 8 and 9.

In addition, the method can provide multiscale modeling capability for arbitrary number of scales. In EXAMPLE 9, the method has been applied to woven composite, which is made of yarn (with microstructure similar to UD composite illustrated in FIG. 7) and matrix materials. The woven microstructure is given in FIG. 12, along with cluster distribution in the matrix and yarn phase. The matrix is modeled with traditional material law, and the yarn is modeled by the Unidirectional (UD) composite microstructure to capture the realistic yarn material microstructure. Since the woven composite is represented by clusters, it can be applied to a higher scale woven laminated woven, depicted in FIG. 7, and realized as a 3-scale composite model.

In one embodiment, the woven microstructure database is used to perform a woven shear simulation, the loading direction is given in FIG. 167. The simulation is a three-scale model, where the macroscopic model is a single element model in a finite element framework (FEF). A 2-scale woven microstructure database, whose yarn phase is modeled as a homogenized anisotropic elastic material, is also tested.

As shown in FIG. 167, the 3-scale woven provides highly nonlinear woven responses due to the yarn elasto-plastic behavior. The 3-scale model is able to capture such non-linearity due to yarn plasticity, which cannot be captured using a simplified 2-scale model with elastic yarn properties. FIG. 167 Woven RVE shear loading simulation by a single element Finite Element model. The results from 3-scale woven model and 2-scale woven model are plotted.

According to the invention, the method can also utilize arbitrary shapes within fixed bounds shown in EXAMPLES 8 and 9. As shown in EXAMPLE 9, the method can be applied to both fiber reinforced composite and woven composite, where the fiber reinforce composite MVE has been illustrated in FIG. 9 and the woven composite MVE is illustrated in FIG. 12.

The method can be applied to arbitrary number of phases/constituents (EXAMPLES 4-5). In certain embodiments, the method has been applied to 2-phase and 3-phase filled rubber composite, as illustrated in EXAMPLE 5 and FIGS. 83 and 86. Referring to FIG. 83, domain decomposition of 2-phase filled rubber into a reduced order model is represented by 64 clusters. Left panel is an original voxel mesh for the 2-phase filled rubber; and right panel is a compressed 2-phase filled rubber model with 32 clusters in the matrix phase and 32 clusters in the filler phase.

FIG. 13 shows a random grain structure with 35 grains (panel (a)), where each grain is considered a different material phase, and grain-by-grain stress predictions, with progressively more clusters per grain (panel (b)), as disclosed in EXAMPLE 4.

The method is predictive (solutions provided outside the bounds of known data), see EXAMPLES 3 and 9. In a certain embodiment, the method is implemented for concurrent modeling for UD Carbon Fiber Reinforced Polymer (CFRP) as described in EXAMPLES 4 and 11. The predictive capability is illustrated in FIGS. 11, 12, 83, 86, and 167, see EXAMPLES 3 and 9.

Additionally, the method is also descriptive (provides homogenized and full field information), as shown in EXAMPLES 3 and 8.

Furthermore, the method could be implemented in hierarchical modeling schemes, see EXAMPLE 6. The method could also be implemented for concurrent modeling schemes, as discussed in EXAMPLES 3 and 8, in which several case studies for concurrent capture of macroscale physical field evolution and microscale physical field evolution are presented, and for combined hierarchical and concurrent modeling schemes (EXAMPLE 7), as explained in FIG. 14, which is an illustration of an implementation with hierarchical modeling between mesoscale and microscale, and concurrent modeling between macroscale and mesoscale. The UD MVE elastic properties are pre-computed and passed to yarn, constituting to a hierarchical modeling process. When the woven MVE is under external loading, its response is computed using both yarn (given by the UD MVE) and matrix properties. When the FE model is under external loading, its local responses is computed using the woven MVE. The FE model and woven MVE responses are computed in a concurrent fashion, establishing a concurrent modeling scheme. A combined hierarchical and concurrent modeling is thus implemented.

Other applications of the method include materials design (material phase selection), as illustrated in FIGS. 4, 9, and 14, where the material properties of each phase will change the MVE behaviors. For example, as shown in FIG. 4, different fiber and matrix properties will result in different stress and strain curves, and the prediction can be made within seconds. This provides the opportunity to generate a large material response database so the optimal material properties of each phase can be selected.

EXAMPLES 3 and 8 disclose embodiments of the invention for microstructure design. In addition, the method can be used with any underlying data representation: images, particles, meshfree, finite element meshes, etc., see EXAMPLES 2 and 5-6. The case shown in FIG. 83 utilizes a voxel mesh directly generated from the 3D TEM process, see EXAMPLE 5.

In one embodiment, the method can be used with machine learning for expedited efficiency (GPU and different NNs), e.g., EXAMPLE 1. The method can also be used to create a microstructure response database that contains many pairs of stress and strain states. The database can then be used as an input to the supervised learning algorithm, one type of machine learning method, to training a Feed Forward Neural Network (FFNN) that can further improve the efficiency. The FFNN can predict stress state given arbitrary strain state that is within the microstructure response database. As shown in Table 1-5 a speed-up of 10000 can be achieved with FFNN.

In certain embodiments, data-driven materials and structure design are achieved with machine learning techniques. As presented in EXAMPLE 1, a microstructure-based optimization is formulated. It allows to perform a topology optimization to design a structure with microstructure damage information, as shown in FIGS. 12-15, 58, 83, 86 and 167.

Process-structure-property-performance for different processes (Additive Manufacturing (AM), composite, polymer matrix composite (PMC)), leading to system design including experiments to predictions can be found in EXAMPLE 6.

In addition, multiscale structure-property materials and structures design are illustrated by two examples (composites and alloys) in FIGS. 4 and 7. As shown in FIG. 7, the method can be used for a 3-scale material system, and can be extended to N-scale.

For composite materials, such as woven composites, the method can be used for fast prediction of the overall material yield surface, as described in EXAMPLE 7. The method can create a ROM for the woven MVE, which provides efficient prediction of woven composite elasto-plastic material responses. As shown in panel (a) of FIG. 15, Table 4 and EXAMPLE 7, yield stresses are computed within one minute, providing considerable savings compared to experimental approaches. The 3D yield surface is shown in panel (b) of FIG. 15, and the yield surface can be used to determine if an applied stress state results in yielding of the material.

In addition, the method can be used to generate a vast woven composite yield surfaces based on different microstructure and material properties of yarn and matrix, providing an enlarged material design space while minimizing computational cost.

TABLE 4 Six yield points used to calibrate the yield surface as shown in EXAMPLE 7 Yield stress components (GPa) σxx σyy σz σyz σxz σxy Yield point 1 0.093 0 0 0 0 0 Yield point 2 0 0.093 0 0 0 0 Yield point 3 0 0 0.076 0 0 0 Yield point 4 0 0 0 0.031 0 0 Yield point 5 0 0 0 0 0.031 0 Yield point 6 0 0 0 0 0 0.041

Without intent to limit the scope of the invention, examples according to the embodiments of the present invention are given below. Note that titles or subtitles may be used in the examples for convenience of a reader, which in no way should limit the scope of the invention. Moreover, certain theories are proposed and disclosed herein; however, in no way they, whether they are right or wrong, should limit the scope of the invention so long as the invention is practiced according to the invention without regard for any particular theory or scheme of action.

Example 1 Clustering Discretization Methods for Generation of Material Performance Databases in Machine Learning and Design Optimization

Mechanical science and engineering can use machine learning. However, data sets have remained relatively scarce; fortunately, known governing equations can supplement these data. This exemplary study summarizes and generalizes three reduced order methods: self-consistent clustering analysis, virtual clustering analysis, and FEM-clustering analysis. These approaches have two-stage structures: unsupervised learning facilitates model complexity reduction and mechanistic equations provide predictions. These predictions define databases appropriate for training neural networks. The feed forward neural network solves forward problems, e.g., replacing constitutive laws or homogenization routines. The convolutional neural network solves inverse problems or is a classifier, e.g., extracting boundary conditions or determining if damage occurs. In this example, we explain how these networks are applied, then provide a practical exercise: topology optimization of a structure (a) with non-linear elastic material behavior and (b) under a microstructural damage constraint. This results in microstructure-sensitive designs with computational effort only slightly more than for a conventional linear elastic analysis.

Computational methods in materials mechanics have evolved with the development of computation tools. A recent advance in computer sciences is the development of the so-called “Big Data” era, where a combined explosion in the number of sensors and datapoints along side computational resources and methods have enabled tracking and using large databases to develop understanding of the world, often replacing smaller more targeted studies that may produce less generalizable results or lack key insights. Taken in the context of computational mechanics, we can develop data-driven computational tools that rely on vast amounts of background data to facilitate, e.g., real-time multiscale simulations for fast multistage material system design, in-the-loop mechanics for controls (e.g., in manufacturing).

In order to develop such data-driven computational tools, two primary areas of study have emerged: (1) generation of materials system databases for materials mechanics, typically using data compression, to reduce computational complexity; and (2) utilization of the database for real-time response prediction and multi-stage design.

The first area arises because data science relies heavily on the size and reliability of the database available. Unlike traditional applications in data science, such as image detection/recognition or automatic control, extremely large and well defined dataset are generally unavailable, or merely inaccessible, in computational mechanics. Whether databases are constructed from experiments or computational models, the cost to generate data at a scale used in, e.g., image recognition has thus far been largely insurmountable. One approach to this challenge has been to use multiscale simulations of material systems using fast calculations of the overall stress of a representative volume element (RVE) for arbitrary far-field deformation loading. Many methods have been developed with the goal of finding an appropriate balance between cost and accuracy for such a problem; these are generally referred to as reduced order methods (ROMs), and many have been developed. In certain embodiments, RVE is a specific case of the MVE which will result in a converged solution of the microstructure it represented. The method can be applied to a RVE, but also to a MVE to cover a wide range of material systems. The second area has often been addressed with methods from machine learning and neural networks to provide real-time prediction and multi-stage design. Data-driven methods have also been used to enhance computational mechanics by, for example, optimizing numerical quadrature and replacing empirical constitutive laws with experiment data. Recently, a deep material network method was proposed which mimics neural networks topologically to link the micro material stiffness to the macro material stiffness. Once trained on a pre-simulated micro-macro stiffness database it can be used to compute, with significant speed-up, the overall stress of an RVE under arbitrary far-field deformation loading.

In this exemplary study, three techniques in modeling microstructure based on data mining are explored and generalized first. These kinds of fast methods address the first area: they can be used to generate the type of very large databases required for the pure or mechanics-enhanced machine learning. The workings of two different classes of neural networks are then derived. Next, an engineering application for these networks, topology optimization considering microstructure, is explored with detailed examples. Then some possible future directions for data-driven computational approaches are outlined to inspire further research in this emerging field within computational mechanics.

Two-Stage Clustering Analysis Methods

The Self-consistent Clustering Analysis (SCA) and its close relatives Virtual Clustering Analysis (VCA) and FEM Clustering Analysis (FCA) are two-stage reduced order modeling approaches including an offline data compression process and an online prediction process. This is concisely illustrated in FIG. 16. In the offline stage, the original high-fidelity RVE represented by voxels or elements is compressed into clusters. In the online stage, macroscopic loading is applied to the reduced order (clustered) RVE. The system of equations describing mechanical response is then solved on only the reduced representation as a boundary value problem. The notation used in this section is summarized in Table 1-1.

TABLE 1-1 Notation used for two-stage clustering analysis methods X Material point X′ Any other material point n Normal to boundary σM Macroscale stress tensor εM Macroscale strain tensor σ(X) Microscale stress tensor ε(X) Microscale strain tensor s Unit eigenstress e Unit eigenstrain {tilde over (Γ)}(X, X′) Green's operator {tilde over (Γ)}0 Fourier transform of the periodic Green's operator I Counting index for clusters J Another counting index for clusters χI(X) Characteristic function for cluster I DIJ Interaction tensor for discretized Lippmann- Schwinger equation B “Interaction tensor” for FCA Ω Material domain ΩI Domain of the Ith cluster (subset of) c1 Volume fraction of the Ith cluster {tilde over (ε)} Reference material strain {tilde over (C)} Reference material stiffness {tilde over (S)} Compliance matrix for reference material A Strain concentration tensor F(X) Deformation gradient at point X F0 Reference deformation gradient A′ Deformation concentration tensor λ0, μ0 Lame's constants of the homogeneous stiffness tensor ξ Fourier point , −1 Forward and inverse fast Fourier transform techniques Δ▪ Incremental form of an arbitrary variable ▪ r1 Residual of the Ith cluster M Jacobiam matrix of r with respect to Δε I4 fourth-order identity tensor CalgJ Tangent stiffness of the material in the Jth cluster NC Number of clusters NF Number of Fourier points NI Number of integration points NE Number of finite elements Ematrix Young's Modulus of the matrix νmatrix Poisson's Ratio of the matrix Einclusion Young's Modulus of the inclusion νinclusion Poisson's Ratio of the inclusion σY,matrix Yield strength of the matrix εp Effective plastic strain (matrix)

To define the boundary value problem, consider a material occupying Ω⊂d. The goal of homogenization is to find the macroscopic constitutive relation between a macroscopic stress

σ M = 1 Ω σ ( X ) dX , X Ω ( 1 - 1 )

and a macroscopic strain

ɛ M = 1 Ω ɛ ( X ) dX , X Ω , ( 1 - 2 )

where |Ω| is the total volume of the region. We define mathematically the RVE problem as

( · σ = 0 , X ɛ = 1 2 ( u + u ) , X Ω σ = σ ( ɛ ; X ) , X Boundary conditions , ( 1 - 3 )

where σ=σ(ε; X) is a general microscale constitutive law. For the homogenization problem, boundary conditions have to be chosen to satisfy the Hill-Mandel condition. In this exemplary study, the periodic boundary conditions are used: ε periodic and σ·n anti-periodic on ∂.

Continuous and Discretized Lippmann-Schwinger Equation

By introducing a reference material with distributed elastic stiffness {tilde over (C)}(X), it has been shown that the RVE problem with periodic boundary conditions is equivalent to the integral equation


ε(X)={tilde over (ε)}(X)−∫{tilde over (Γ)}(X,X′):(σ(X′)−{tilde over (C)}(X′):ε(X′))dX′  (1-4)

where {tilde over (ε)}(X) is the strain in the reference material when applying the same loading and boundary conditions as the original RVE problem; σ(X′)−{tilde over (ε)}(X):ε(X′) is the eigenstress applied to the reference material; {tilde over (Γ)}(X,X′) is the Green's operator associated with the reference material. The physical meaning of −{tilde over (Γ)}ijkl(X,X′) is the strain component εij at material point X if the unit eigenstress skl is applied at material point X′, with the components of skl defined by smnklkmδln, where δkm and δln are Kronecker delta functions.

The integral equation given in Eq. (1-4) is known as the Lippmann-Schwinger equation, commonly seen to describe particle scattering in quantum mechanics. There is typically no explicit form of {tilde over (Γ)}(X,X′), unless the reference material is homogeneous.

The domain can be decomposed into several sub-regions, called clusters, distinguished mathematically by the characteristic function x as shown in Eq. (1-5).

χ I ( X ) = ( 1 , X Ω I 0 , otherwise ( 1 - 5 )

where I=1, 2, 3, NC denotes each cluster, and ΩI is the subset of the volume within cluster I. These clusters are defined during the offline stage. This allows one to discretize the Lippmann-Schwinger equation, as:


εI={tilde over (ε)}I−ΣJ=1NCDIJ:(σJ−{tilde over (C)}JJ),∀I∈{1, . . . ,NC}  (1-6)

where

I = 1 c I Ω χ I ( X ) ( X ) dX

denotes the volume average of an arbitrary variable ▪ in the Ith cluster;

c I = I Ω

is the volume fraction of the Ith cluster where the volume of the Ith cluster is given by |ΩI|. DIJ is the interaction tensor given by

D IJ = 1 c I Ω χ I ( X ) χ J ( X ) Γ ~ ( X , X ) dXdX . ( 1 - 7 )

The physical meaning of −(DijklIJ) is the average strain component ij in the Ith cluster if the uniform unit eigenstress component kl is applied in the Jth cluster.

Remark 1: If homogeneous reference material is used, as in SCA and VCA, {tilde over (ε)}IM.

Remark 2: A counterpart of Eq. (1-6), similar to the idea of FCA, can be expressed as


σI={tilde over (σ)}I−ΣJ=1NCBIJ:(εJ−{tilde over (S)}JJ),  (1-8)

where {tilde over (S)} is the compliance matrix of the reference material; a is the stress in the reference material when applying the same loading and boundary conditions as the original RVE problem. εJ−{tilde over (S)}JJ is the volume average eigenstrain in the Jth cluster. The physical meaning of −BijklIJ is the average stress component ij in the Ith cluster if the uniform unit eigenstrain component kl is applied in the Jth cluster. Note that the reference material for FCA is the elastic state of the original RVE, instead of a homogeneous reference material as used by the other two methods here.

Remark 3: The relationship between BIJ and DIJ is given by


{tilde over (S)}I:BIJ=−DIJ:{tilde over (C)}J.  (1-9)

This can be proven by noting that the effect of applying any unit eigenstrain kl in some cluster J of the reference material is equivalent to that of applying eigenstress −{tilde over (C)}:ekl in the same cluster.

In the online stage, SCA solves the incremental, discretized Lippmann-Schwinger equation, Eq. (1-6), with arbitrary external loading conditions. These can either be of the fixed strain increment εM type or be of the fixed stress increment σM type.

Offline: Clustering and Interaction Tensor

The offline stage includes three primary steps: (1) data collection, (2) unsupervised learning (e.g., clustering), and (3) pre-computation of the interaction of clusters. Computation of the linear elastic response (data collection) and subsequent clustering based on that response are conducted identically for each of the three two-stage methods presented below: the same voxel mesh and clusters are used in the examples for all three methods. The difference between methods comes in the computation of the interaction tensor.

(1) Data Collection

Data collection provides information used to construct the reduced representation of the system. It typically involves some computation of the response of a fully resolved system, perhaps with a simplified material model, over a limited set of loading cases. One measure of mechanical response that could be collected is the strain concentration tensor used in micromechanics A, defined by ε(X)=A(X):εM, which maps between the far-field or applied strain, εM and the strain measured at point X in the domain, ε(X).

The choice of the strain concentration tensor as given above is often suitable, but it depends on the relevant details of the problem. For example, at finite deformations one might consider using the deformation concentration tensor:

A ( X ) = F ( X ) F 0

where F(X) is the deformation gradient at any given point X within the domain and F0 is the macroscopic deformation corresponding to the boundary conditions. Alternatively, if the elastic and plastic material responses differ substantially, e.g., if one is isotropic and the other anisotropic, including information about the plastic part of the deformation might be desirable.

(2) Clustering

The goal of clustering is to reduce the number of degrees of freedom required to represent the system while minimizing the loss of information about the mechanical response. One way of doing this is by grouping material points within the domain of interest. If one can assume that the material response within each group is identical, the evolution of the domain can be determined by solving for the response of each group rather than each material point.

Using the data generated during the collection phase, one of the many clustering (or unsupervised learning) techniques might be applied to optimize the domain decomposition. In the following examples, Self-Organizing Maps (SOMs) are employed, as illustrated in FIG. 16, where eight clusters (four in each phase) are constructed. The clustering process assigns each material point with a cluster ID, such that clusters are labelled 1, 2, Nc.

The k-means clustering method has also been used. More elaborate clustering schemes might also be considered. Ongoing efforts include “adaptive clustering” schemes that mimic adaptive FE methods in their ability to evolve as deformation progresses, and “enriched” machine learning, whereby a priori information from mechanics about the deformation fields outside the bounds of the data collected in Step (1) are used to guide or bound the unsupervised learning. Another unexplored future direction might use a “feature-based” machine learning that includes microstructure information in addition to mechanics information.

(3) Interaction Tensor

The interaction tensor describes the impact each cluster has on each of the other clusters. Once the clustering process is completed, the interaction tensor can be explicitly computed. Importantly, the integral part only has to be computed once during the offline stage. Only the results of that calculation are then used for the online stage. Three ways to compute the interaction tensor, one used by each of the methods highlighted here (though these are not exclusive to each), are:

A. SCA: Fourier Transform for DIJ

With periodic boundary conditions and homogeneous reference material, the Green's operator has a simple expression in Fourier space, given by

Γ ^ i j k l 0 ( ξ ) = δ i k ξ j ξ l 2 μ 0 ξ 2 - λ 0 2 μ 0 ( λ 0 + 2 μ 0 ) ξ i ξ j ξ k ξ l ξ 4 ( 1 - 10 )

where {circumflex over (Γ)}ijkl0=(Γ0) is the Fourier transform of a periodic Green's operator Γ0; λ0 and μ0 are the Lamé's constants of the homogeneous stiffness tensor; is the Fourier point. Then the interaction tensor can be calculated with

D IJ = 1 c I Ω χ I ( X ) - 1 ( ( χ J ) ( Γ 0 ) ) d X , I , J { 1 , , N C } ( 1 - 11 )

using the fast Fourier transform (FFT) technique. The computational complexity is O((NC)2(NF)log(NF)), where NF is the number of Fourier points used in the FFT calculation.

B. VCA: Numerical Integration for DIJ

With an infinite homogeneous reference material, the Green's operator can be expressed in real space. Numerical integration is the most straightforward method to compute the integral equation given in Eq. (1-7). The computational complexity is O((N)2), where NI is the number of integration points used.

C. FCA: Finite Element Method for BIJ

Based on the physical interpretation of the interaction tensor, the finite element method can also be used. By applying uniform unit eigenstrain component kl in the Jth cluster, the average stress can be computed for all clusters, resulting in BijklIJ for all I=1, . . . , NC. Thus, the computational complexity is O(6(NC)(NE)), where NE is the number of finite elements used. The tensor BIJ is similar to the interaction tensor DIJ, although BIJ is determined by applying strains rather than stresses.

Online: Reduced Order Response Prediction

Once the interaction tensor database is prepared, the discretized Lippmann-Schwinger equation defined in Eq. (1-6) is solved in the online stage. An incremental form of Eq. (1-6) is given by


εI={tilde over (ε)}I−ΣJ=1NCDIJ:(σJ:−{tilde over (C)}JJ),∀I∈{1, . . . ,NC},  (1-12)

where {tilde over (ε)}I is the applied incremental reference strain. The incremental stress σI is a function of the incremental strain εI according to the local material constitutive laws. So the unknowns of Eq. (1-12) are the strains in each cluster {ε}={ε1, . . . , εNC}. For nonlinear microscale constitutive laws, Eq. (1-12) is nonlinear and has to be solved iteratively. Newton's iterative method is used by SCA and VCA, while a different iterative method is used by FCA. The residual form of Eq. (1-12) is given by

r I = ɛ I - ɛ ~ I + J = 1 N C D IJ : ( σ J - C ~ J : ɛ J ) , I { 1 , , N C } ( 1 - 13 )

Then the Jacobian matrix

{ M } = { r } { ɛ }

for the Newton's method is given by

M IJ = r I ɛ J = δ IJ I 4 + D IJ : ( C a l g J - C ~ ) , I , J { 1 , , N C } ( 1 - 14 )

where I4 is the fourth-order identity tensor. The tangent stiffness of the material in the Jth cluster is CalgJ. An alternative way to solve for the local mechanical responses is to minimize the complementary energy of the clustering-based system.

Comparison of the Methods

In order to compare the accuracy and efficiency of each method, a 2D plane strain model for a two-phase material is constructed and shown in FIG. 17. The 2D mesh contains 600×600 square pixels. The inclusion area fraction is 51%. Material constants of the matrix and the inclusion are given in Table 1-2. The yield surface is von Mises surface as shown in Eq. (1-15). The hardening law of the matrix material is given in Eq. (1-16). Note that this will be approximated by a non-linear elastic behavior for monotonic loading in future sections.

TABLE 1-2 Material constants for matrix and inclusion Ematrix Einclusion Area fraction (MPa) νmatrix (MPa) νinclusion of inclusion 100.0 0.30 500.0 0.19 0.51

f = σ _ - σ Y , m a t r i x ( ɛ _ p ) 0 ( 1 - 15 ) σ Y , matrix = ( 0.50 + 5 ɛ _ p 0 < ɛ _ p 0.04 0.62 + 2 ɛ _ p 0.04 < ɛ _ p ( 1 - 16 )

The equivalent von Mises stress is σ. The yield stress σY,matrix is given by the hardening law in Eq. (1-16) with equivalent plastic strain εp. Strain from 0 to 0.05 is prescribed in the x-direction and zero strain is enforced in the y- and xy-directions. A direct numerical simulation (DNS) of the microstructure under these loading conditions is performed using the FEM and the effective von Mises stress is recorded for later comparison to the reduced order model results.

A one-time data-compression of the microstructure is performed using the strain concentration tensor. The resulting clustering of the microstructure is shown in FIG. 18. The interaction tensor for each method is computed using the aforementioned algorithms.

The different methods, used to compute the interaction tensors, each results in a slightly different form of the tensor. There are strong similarities—after all, the same microstructure and clustering is used—though the details differ. FIG. 19 shows a magnitude plot of each of the three methods, where magnitude represents the effect of each stress component in the Jth cluster on the corresponding strain component in the Ith cluster. FIG. 19 shows component-wise magnitude plots for DSCAIJ, DVCAIJ, BFCAIJ in panels (a)-(c), respectively. Spikes along the diagonal direction for all three interaction tensor surface plots suggest self interaction has more contribution than the rest of clusters in cluster-wise stress increment. DSCAIJ and DVCAIJ have the similar magnitude along their diagonal direction due to the homogeneous reference material assumption. BFCAIJ has different magnitudes for matrix and inclusion phase along the diagonal direction, implicitly representing a heterogeneous reference material. FIG. 20 is plots for DSCAIJ, DVCAIJ, BFCAIJ in profile; note that for FCA the two regions correspond to different physical domains (matrix and inclusion). The trends of the magnitudes shown in FIG. 19 and FIG. 20 suggest that inter-cluster interaction is not as strong for VCA as for SCA and FCA, although this difference is relatively minor. The strongest interactions are intra-cluster, as shown by the peaks along the diagonal. FCA has two distinct regions of peaks, corresponding to the set of clusters in the inclusion and in the matrix. The tensor is constructed in an ordered way, which results in these two distinct sets of clusters. This is unlikely to change the overall solution accuracy.

Once the microstructural database is created and the interaction tensor has been computed, the online prediction is performed. FIG. 21 shows that the ROM results for stress in the x-direction in all three cases are within 5% of the DNS results, with small differences between the three ROMs. Panel (b) of FIG. 21 highlights the slight differences between the three methods: SCA follows the same trend as the DNS, but is slightly softer; the response predicted by VCA is softer still, as a result of the constant reference material assumption; the trend exhibited by FCA is slightly different from the DNS overall, although still quite close in value. FIG. 22, σyyM plotted against εxxM, shows that SCA has the best agreement with the DNS solution, and is the only prediction within 5% of the reference solution. Note that VCA has different boundary conditions than the DNS solution, i.e., it uses a fictitious surrounding domain. Some deviation from a DNS result with periodic boundary conditions is therefore expected.

Machine Learning on Databases Generated with Predictive ROMs

Neural networks are a specific class of machine learning algorithms, which in the most basic form appear similar to regression analysis. In practice, these methods involve modifying input data through a series of functions to obtain output data. The exact series of functions and their weights and forms depend on the application. These methods can provide an increase in speed over more conventional approaches, once the algorithm has been appropriately trained. In order to have a highly effective modeling approach based on machine learning, rich databases of mechanical response information are required to perform that training. Developing such a database with experiments is intractable, particularly for design of new materials or material systems where no material performance information exists. The fast, predictive models (SCA, VCA, FCA) outlined above are thus desirable for quickly populating relatively large materials databases. This enables the use of machine learning algorithms in multiscale design, where simultaneously application and satisfaction of criteria and constraints governing the material microstructure and component-level macrostructure is required.

Feed forward neural networks (FFNNs) were the first neural networks developed. The FFNN was designed to learn complex input-output relations. As such, FFNNs can be used to replace conventional constitutive laws; this is particularly appealing when descriptions of the homogenized behavior of a material is complex and/or difficult to obtain. The basic structure of an FFNN includes an input layer, hidden layers, and an output layer. Every pair of neurons in neighboring layers have a weighted connection. Each neuron in hidden layer and output layer has a bias. In FFNNs, neurons in the same layer are not connected. In the learning procedure, the connection weights are changed following a predefined set of rules, such as with back propagation. Funahashi and Hornik et al. proved that three hidden layers in an FFNN is sufficient to learn any non-linear continuous function.

Early efforts to apply machine learning to mechanics used a computation and knowledge representation paradigm, which is actually a type of feed forward neural network, to directly “learn” material behavior by training from analytic and experimental data. One early work applied a back-propagation neural network to model the behavior of concrete in plane stress under monotonic biaxial loading and compressive uniaxial cyclic loading. Once an RVE model is established, for example, an FFNN can be trained on that data, and a concurrent multiscale scheme to directly connect microstructure to the macro-scale material response might be achieved, with the FFNN replacing the RVE or constitutive law to describe the response at each material point.

Neural networks have been studied with the goal of integration with multiscale methods. In such cases, some parts of numerical simulations are replaced with neural networks to better utilize their merits. With the development of numerical methods and the increasing interest in multiscale simulation, integration of multiscale simulation and neural networks is continuously developing.

In design optimization, one might expect there to be many calls to the material subroutine (e.g., one for each element for each load for each design iteration). Thus running ROMs for RVEs might still be time consuming, compared to a model predicting RVE stress responses given a strain state within milliseconds, or a model providing strain state given microstructure stress contours within milliseconds. To tackle this issue, we propose replacing the ROMs by neural networks trained on the RVE responses computed with SCA for micro-stress and macro-stress. These networks preserve microstructure information and are engineered to:

    • Predict macro (homogenized) stress or micro (local) stress given a macro-strain using an FFNN for one RVE. In this case, the FFNN plays the role of traditional material constitutive equations and homogenization (to compute the macro-stress). We denote these FFFNmicro and FFNN, respectively
    • Predict macro-strain given any micro-stress distributions using a convolutional neural network (CNN). This is the inverse of the FFNN. The input is a stress distribution within an RVE domain. The output is the macro strain loading applied to this RVE (the boundary conditions). We denote this CNN.
    • Compute RVE damage by comparing local von Mises stress to a stress threshold. A CNN is trained to identify the onset of damage within an RVE. In this case, the CNN acts as a classifier, which identifies whether or not the applied macro-strain will cause microstructural damage. We denote this CNNclassify.

Database Generation for Machine Learning Using SCA

The ROMs presented in this example can be used to generate a database for machine learning. In this case, we are using SCA with 8 clusters. In general, such databases contain NT training samples and NV validation samples. For the RVE model, a database with NT=1000 strain-stress pairs is computed for a nonlinear elastic material by randomly sampling 200 terminal states with four sub-loading steps each, as shown in FIG. 23. Similarly, NV=150 strain-stress pairs (30 final states, plus four intermediate steps) are generated for validation. A monotonically increasing load is assumed and all the final points are confined to a spherical space with a radius of 0.05. Table 1-3 shows the reduction in time required to build a database achieved by using SCA, versus FEM or FFT for this example case. Without a reduced order model, the generation of microstructure database takes days using FEM or FFT. The database contains NT=1000 pairs of macroscopic strains and stresses εM,s, σM,s, and local (micro)stress σs(X), where s=1, 2, . . . , NT. In this case, we only consider a plane strain problem and do not consider stress in the z direction. SCA reduces the total computational time by two orders of magnitude compared to FFT and by four orders of magnitude compared to FEM.

FIG. 23 shows random samples of strain state; two hundred final states were selected, and four evenly spaced intermediate steps to reach the final states were recorded for a total of NT=1,000 samples. All strain states will be applied to the RVE to generate corresponding stress states.

TABLE 1-3 Comparison of time required to generate a microstructure database for nonlinear elastic RVE responses with NT = 1,000 samples of strain states. SCA requires only one single workstation to generate the database. Speedup over Method Total time (s) FEM FEM 2.04 × 107 FFT 3.01 × 105 68 SCA offline: 13.0 + online: 9 400 (4 + 4 clusters) 2.16 × 103

Feed Forward Neural Networks

In order to illustrate the structure of feed forward neural networks (FFNNs), a simplified one dimensional example is presented. In linear elasticity, stress is related to strain by the material stiffness; this can be generically defined as a mapping. The overall structure of a neural network can also be described as a mapping, i.e.:

( Constitutive equation : Stress = C o n s t i t utive ( strain ) Neural network mapping : Stress = F FNN ( strain ) ( 1 - 17 )

where FFNN is the FFNN that uses strain state ε as input, and generates stress state σ as the output. The structure of a simple FFNN is shown in panel (a) of FIG. 24. As shown in FIG. 24, panel (a) is illustration of an FFNN network with one hidden layer for a linear elastic example; the collective function of the weights and biases connecting the input layer (green), the hidden layer (blue), and the output layer (red) is that of Young's Modulus E. panel (b) is s stress-strain diagram, showing how the input strain is interpreted by the FFNN for a linear elastic case using linear activation functions and zero biases. The notation used in this figure and throughout this section is defined in Table 1-4. For this illustration case, only one sample is considered, hence s=1 and all variables are written without the superscript s. A general FFNN contains neurons (the circles) and weights (black lines). In general, an FFNN has one input layer, one output layer, and multiple hidden layers. Each layer may have multiple neurons; for the input and output layers, these are simple the input and output values. In the simplest case, an FFNN would have one input neuron, one hidden neuron, and one output neuron. For 1D linear elastic stress analysis in such a case, the input neuron would be strain, the hidden neuron would act as a multiplicative, functional decomposition of the stiffness that recovers the total stiffness required to map strains to stress in the output neuron. Generalizing this slightly, we might consider an FFNN with three hidden neurons, as shown in FIG. 24. Each neuron has only one value. The first neuron is simply the strain:


ai=1l=1=ε(input layer)  (1-18)

The three neurons in the hidden layer take in this value, and each take on the value given by:


aj=1,2,3l=2=(Σi=11Wijl=2ail=1+bjl=2)(hidden layer)  (1-19)

where is an activation function. In the training part, this example uses Sigmoid function:

f ( x ) = 1 1 + e - x ,

and each neuron is computed using a different weight Wijl=2 and bias bjl=2, where i is the neuron in the previous layer (in this case, the input later) and j is the neuron in the current layer (in this case, the hidden layer). Finally, the overall response—the stress—is given by:


σpredicted=ak=1l=3j=13Wjkl=3ajl=2+bkl=3 (output layer)  (1-20)

The combination of all the W and b terms, as well as the activation function, work as the fitting factors in a regression analysis. The activation function is fixed for all neurons, and is used to condition the weighting factors. For the constitutive model outlined above, the overall results of the weights, bias and activation function would perfectly match the elastic modulus. If we only consider weights and the activation function is just a linear mapping, the function of each neuron in hidden layer is given explicitly in panel (b) of FIG. 24, where the bias is taken as zero for all neurons.

The physical interpretation of individual neurons is more complicated for non-linear responses, although the overall idea is the same. In this formulation, strain path dependence (i.e., plasticity) is impossible to capture, and the net result is a response map where one might think of the collection of neurons (weights, biases and activation functions) as an “instantaneous elastic modulus,” or the slope of a line that relates strain to stress at the current point in strain.

FIG. 25 is an illustration of an FFNN with multiple hidden layers; NL: index of layers, NN(l): number of neurons in layer l. The formulation of the FFNN is given in Eq. (9-21) with associated interpretation of the FFNN structure. The indices i and j represent the neuron ID in the previous layer and current layer, e.g., W12l=2 is the weight between neuron 1 in layer l=1 and neuron 2 in layer l=2. The example shown in FIG. 25 extends the previous example to consider a two-dimensional stress analysis, with many neurons per layer and several layers. The inputs in are three macroscale strain components εxxM, εyyM, γxyM, in a plane strain problem. The outputs are the three macroscale stress components εxxM, σyyM, τxyM. The stress component σzzM is not considered. The samples (stress-strain pairs) from the database outlined above are used to train the neural network. After training, the FFNN can predict stresses when given strain inputs.

In each layer of a general FFNN, each neuron takes the output value from each neuron in the preceding layer as inputs and gives a single output. This is repeated for each layer. Generalizing Eqs. (1-18), (1-19), and (1-20) to an arbitrary number of layers and neurons per layer results in Eq. (1-21), where the value of the jth neuron in layer l for the sth sample (either a training sample or prediction) can be expressed as:

( 1 - 21 ) a j l , s = ( ɛ j M , s , ifl = 1 ( input layer ) 𝒜 ( i = 1 N N ( l - 1 ) W ij l a i l - 1 , s + b j l ) , ifl { 2 , , N L - 1 } ( hidden layers ) i = 1 N N ( l - 1 ) W ij l a i l - 1 , s + b j l , ifl = N L ( outout layer )

where the final layer gives the estimated stress:


σprediceted,jM,s=ajNL,s.  (1-22)

TABLE 1-4 Notation table of variables used in the feed forward neural network εjM,s Macroscale strain tensor, s = 1, . . . , NT s Counting index for number of samples (training or validation, depending on context) l Counting index for number of layers i Counting index for neurons in a given layer j Counting index for neurons in another layer NT Number of training samples NL Number of layers in the neural network NN(l) Number of neurons in layer l Wijl Weight connecting the ith neuron in layer l − 1 to the jth in layer l bjl Bias of the jth neuron in layer l ajl,s Neuron value for jth neuron in lth layer and for sth sample Activation function FFNN Feedforward neural network function

The outputs ajl,s of each layer possesses similar physical meaning as explained for the 1D case. The input strain components are represented by l=1, ajl,s for the sth sample, and l=2, . . . , NL−1, ajl,s represents an estimate of the the nonlinear stress responses of the microstructure. The activation functions and weights of layer l=2, . . . , NL−1 play a roughly similar role to the classic definition of the tangent modulus in solid mechanics. The hidden layers take in strain components and produce an estimate of the non-linear stress responses. During the training process, the non-linear relationship between stress and strain is gradually “learned” by those hidden layers. In the last layer, l=NL, ajl,s represents the predicted stress components. The predicted stress components are produced through the regression operation in the output layer, as shown in Eq. (1-21). The weights and bias of the output layer correct the prediction generated from hidden layers, and produce accurate nonlinear stress responses. In order to make the concept clear, one might consider the hidden layer as unitless values operating on intermediate strain values, while the units of W and b in output layer are those of stress (e.g., MPa in the example problem). The FFNN can learn nonlinear elastic material behaviors due to following two key factors: 1) hidden layers approximate the material nonlinear elastic responses as a traditional constitutive model would do, as described in Eq. (1-17); 2) the output layer corrects the predicted nonlinear responses for improved prediction accuracy.

Feed Forward Neural Network with Database Generated by SCA

In this case study, an FFNN is trained with data generated using SCA. Using the same microstructure as given in FIG. 18, SCA computes the stress responses of the RVE (or any arbitrary RVEs) when a monotonically increasing strain is applied. Since SCA provides efficient evaluation of the stress state, it is convenient to train the FFNN on data made with SCA. The FFNN can then replace SCA by “learning” the stress state as a function of the strain state. This would establish a straight-forward relationship between strain and stress for near-instantaneous evaluation of RVE stress responses. Note that although a plastic material is described for the matrix material in the RVE analyzed by SCA, the FFNN described here is limited to non-linear elastic (i.e., path independent) material behavior: we approximate the plastic response with a non-linear elastic one, and focus on monotonic loading. Moreover, in the following design case, the FFNN is trained to predict not only the overall RVE stress responses (as was shown in FIG. 25), but also cluster-wise, local stress responses. This is essentially the same process, but the output layer is size NC×NN (l=1), with one point for each stress component for each cluster. This second form replicates the non-homogenized results of SCA.

Training

The training procedure for an FFNN can be reformulated as an optimization problem. We define the loss function (or cost function) as Mean Square Error (MSE) for the estimated stress and stress computed by RVE using SCA. Assuming one hidden layer, the optimization formulation is given by:

find : W ij l = 2 , b j l = 2 , W jk l = 3 , b k l = 3 minimize loss function : MSE = 1 N T × N N ( l = 3 ) s = 1 N T k = 1 N N ( l = 3 ) ( σ k l = 3 , s - σ k * , l = 3 , s ) 2 where : σ k l = 3 , s = j = 1 N N ( l = 2 ) W jk l = 3 ( 𝒜 ( i = 1 N N ( l = 1 ) W jk l = 3 ɛ i M , s + b j l = 2 ) + b k l = 3 ) ( 1 - 23 )

By finding the optimal values for Wijl=2, bjl=2, Wjkl=3, and bkl=3, MSE is reduced. Note that only training data is used in this process, hence s=1, 2, . . . , NT.

Usually, the MSE gradually decreases with each training step. To ensure the trained neural network is general enough for all possible input states, some data points called verification data are used to monitor trends in the error. The minimization iterations terminates before the error of the verification data starts to increase. This ensures the neural network is able to provide certain extrapolating capability for data points that are not within the training set.

The FFNN described above was trained on the database. In this case, an FFNN with one hidden layer and 50 neurons was chosen. In the training procedure, 1,000 samples are used to train the neural network. The Levenberg-Marquardt optimization algorithm is used to reduce the MSE.

Prediction

After the training process, a fast evaluation of the stress state during a monotonic loading process was performed. The FFNN used for this predicts the macroscale stress tensor for a given macroscale strain tensor following Eq. (1-24); similarly, Eq. (1-25) is used to predict the local (cluster-wise) stress tensors given a macroscale strain tensor.


σM,s(FFNN)kl=NL,s=FNNiM,s)  (1-24)


σs(FFNN)(X)=σkl=NL,s(X)=FFNNmicroiM,s)  (1-25)

To demonstrate validation of the macroscale FFNN, the l2 norm is used to measure the difference between the overall stress predicted by Eq. (1-24) and the homogenized SCA results for each sample in the validation data set, as computed by:


DifferenceFFNNs=∥FFNNiM,s)−σM,s(SCA)2·,s=1,2, . . . ,NV,  (1-26)

To validate the trained FFNN, another 30 final strain states (unknown during the training) with five load steps each (including the final state) were selected. FIG. 26 shows a histogram of difference measured with the l2-norm, Eq. (1-26), for these new NV=150 strain-stress pairs of in the validation data set. Most of the test samples have a very low l2-norm, which shows that the FFNN is well trained. FIG. 27 shows the SCA stress data for each stress component plotted against the FFNN predictions; a perfect match has a slope of one. The associated cross-correlation statistic is one: the FFNN solutions match the SCA solution perfectly. At each load step the stress predictions of the FFNN and SCA match, as shown in FIG. 28.

This case study illustrates a convenient workflow that used the reduced order modeling approach to generate a rich microstructure response database for training an FFNN, which is then used for generating fast predictions of the RVE responses. Note that although the validation for the FFNN for the homogenized stress-strain relationship is given in detail here, a similar process has been used for the relationship between macroscale loading and microscale (cluster-wise, or local) stresses, as given in Eq. (1-25).

We propose that the FFNNs shown here can be used in a design optimization process, such as topology optimization or microstructural design, where a fast and accurate material responses prediction is desired. However, note that the material is non-linear elastic and/or under monotonic loading. If plasticity and loading/unloading are considered, a different FFNN setup or a different neural network may be required. A speed comparison of running 150 samples with SCA and FFNN is given in Table 1-5, where the speedup for online prediction of σM is 10000 for the FFNN over SCA. This idea will be explored further using both the FFNN predictions (Eq. (1-24) and Eq. (1-25)) and convolutional neural networks.

TABLE 1-5 Comparison of time required to run 150 samples using SCA and FFNN. Online speedup Method Total time (s) over SCA SCA (4 + 4 offline: 13.0 + clusters) online: 3 × 102 FFNN training: 60 + 10 000 prediction: 3 × 10−2

Convolutional Neural Network

Convolutional networks or convolutional neural networks (CNNs) are widely used in fields such as image recognition and feature identification. The term “convolutional” refers to the linear mathematical operation and indicates that the convolution operation is implemented in at least one layer of the network rather than conventional matrix multiplication. This is a biologically inspired model used to handle known grid-like topology data such as time series (1D grid of samples at successive time intervals) or image data (2D grid of pixels). Convolution neural networks have been implemented in material science and multiscale modeling to analyze the microstructure properties where the input data are microstructure images. Extracting material information through microstructure images, a ubiquitous data type in materials science, has proven to be a promising application of CNNs. For example, Lubbers et al. implemented a CNN based on the distribution of texture images for unsupervised detection of low-dimensional structures. Scanning electron microscope (SEM) images are frequently used in materials science to distinguish between categories of materials. Such image datasets can be classified with a single feature or with multiply features using CNNs. Some studies have implemented CNN to featurize SEM images over a single set of data. However, a scalable and a generalizable feature should be used to facilitate widespread applicability of the CNN. Ling et al. analyzed the generalizability and interpretability of CNN-based featurization methods for SEM images, and found that mean texture featurization is generally useful in such cases, although sometimes feature-free CNN procedures are appealing as well.

The application of CNNs is not limited to images. Cang et al. established a CNN approach to predict the physical properties of a heterogeneous material, replacing standard statistical or micromechanical modeling techniques. The generated scheme is applicable to systems with a highly non-linear mapping based on high dimensional microstructure. They have implemented a convolutional network to quantify material morphology followed by another convolution network to predict the material properties given the microstructure. This can also be done in 3D, for complex materials and responses.

A CNN model includes several basic unit operations: padding, convolution, pooling, and a feed forward neural network (FFNN). The structure of an example 1D CNN is shown in FIG. 29.

The input is a series of stress values, i.e., a 1D problem. The 1D CNN includes several loops of padding, convolution, and pooling. For a specific loop iteration η, a padding procedure adds zeros around boundaries, to ensure that the post-convolution dimension is the same as the input dimension. After padding, several kernel functions will be used to approximate the discrete convolution operator given by:


{tilde over (σ)}xκ,ηξ=−(Lconv−1)/2(Lconv−1)/2wξκ,ησx+ξpadded,η+bκ,η,  (1-27)

where σx+ξpadded,η is the input, wξκ,η is the κth kernel function, and bκ,η is the bias, for ηth convolution process and η=1, 2, . . . , Nconv. The size of the kernel function is Lconv. A summary of all of the notation used in this section is given in Table 1-6. The convolution operation can be regarded as a feature identification operation. After padding, a pooling layer will decrease the dimension of inputs, and extract the most important features from the post-convolution data. A one dimensional max pooling equation is given by


{circumflex over (σ)}αmax,κ,η=MAX({tilde over (σ)}ξκ,η,ξ∈[(α−1)Lpooling+1,αLpooling])


α=1,2, . . . ,Npoolingη  (1-28)

where {circumflex over (σ)}max,κ,η is the output value, {tilde over (σ)}ξκ,η is the input value, and Lpooling is the length of the pooling window. Max pooling extracts the maximal value from the window, but other pooling operations might also be used. In this case we surmise that, to predict remote strains, the maximum stress values might be telling. Padding, convolution, and pooling may be repeated for Nconv times. The value will be transferred to a fully connected FFNN, such as that illustrated in the previous section.

FIG. 30 represents a generalized overview of the structure of a typical two dimensional CNN. Continuing with the two-dimensional example problem given above, for a CNN the input is the stress within the 600×600 grid, given by σ(α,β) were α and β correspond to the x- and y-components of the stress map. At each grid point (α,β), the three stress components xx, yy and xy are stored. Padding padding will add zero boundaries around the input data to make sure that after convolution, the size of the maps will remain same.

Similar to 1D convolution, the convolution operation applies a kernel function over the stress contours and generates a new feature map with the same resolution as the initial stress contours. The extension of the convolution operation to two dimensions for ηth convolution process is shown in Eq. (1-29).


{tilde over (σ)}α,βκ,ηζ=1Nfeatureξ=−(Xconv−1)/2(Xconv−1)/2Σψ=−(Yconv−1)/2(Yconv−1)/2Wξ,ψζ,κ,ησα+ξ,β+ψpadded,ζ,η)+bκ,η  (1-29)

where κ is the kernel ID of the convolution layer and goes from one to Nkernel. The size of the convolution kernel in dimension 1 and 2 are given by Xconv and Yconv, and are both odd numbers. The counting indices in the kernel in dimension 1 and 2 are defined ξ as and ψ, respectively. The number of stress components is Nfeature; for this 2D example, Nfeature=3. By applying the kernel to each element in each the input stress array, a complete feature map will be generated.

To define a nonlinear relationship between the input and output using the CNN, a nonlinear activation function is often used. In some cases, this is a Rectified Linear Unit (ReLU) layer, which is applied to all feature maps generated from the convolution operation. For simplicity of illustration, this step is not shown in the equations and figures.

A pooling layer is applied to all feature maps after the ReLU layer to compress the resolution of the data in the X and Y directions. Different pooling operations might be used; in this example, we selected max pooling. The max pooling operator divides the feature map into many subset regions, and selects the maximum value from each region to use as the value in the compressed feature map; generalizing from Eq. (1-28), for the nth convolution process, this can be written as:


{circumflex over (σ)}α,βmax,κ,η=MAX({tilde over (σ)}ξ,ψκ,ηξ∈[(α−1)Xpooling+1,αXpooling],ψ∈[(β−1)Ypooling+1,βYpooling])


α=1,2, . . . ,β=1,2, . . . ,NYpoolingη  (1-30)

After the final pooling operation, all compressed feature maps are converted into a single vector through a flattening operation. The flattened array is then used as the input of the FFNN for regression to compute the corresponding strain. Further details of the CNN method and implementation can be found in literature cited in the beginning of this section.

TABLE 1-6 Notation used to describe the CNNs shown above. X A point in the microscale (inside the RVE) region σM Macroscale stress tensor εM Macroscale strain tensor σ(X) Microscale stress tensor ε(X) Microscale strain tensor α First direction in X β Second direction in X {tilde over (σ)} Microscale stress tensor to which a kernel has been applied {circumflex over (σ)} Microscale stress tensor to which a kernel and pooling has been applied NT Number of training samples in the database NV Number of verification samples in the database Nkernel Number of kernels in convolution Npooling, NXpoo Size of output after pooling Nfl Number of entries in flattened vector Nconv Number of repeats of padding, convolution, and pooling in CNN Wijl Weight in FFNN connecting the ith neuron in layer l − 1 to the jth in layer l bji Bias in FFNN of the Jth neuron in layer 1 ξ Counting index for location within kernel in dimension 1 ψ Counting index for location within kernel in dimension 2 ζ Counting index for features κ Counting index for kernels η Counting index for convolutions Wξ,ψζ,κ,η Weight connecting the input for ζ feature and κ kernel on convolution layer bκ,η Bias of the kernel κ on the convolution layer Lconv, Xconv, Size of the kernel Yconv Lpooling, Size of the pooling window Xpooling, Ypooling padding Padding function conv Convolution function pooling Pooling function pcp Combined padding, convolution, and pooling operation flatten Flattening function CNN Convolutional neural network function CNNclassify Classification convolutional neural network function d Binary indicator given by classification CNN

Convolutional Neural Network for Boundary Condition Identification with Database Generated by SCA

Using the CNN illustrated in FIG. 30, a mapping has been established between local stress distribution and applied external strain on the microstructure.

Training

This CNN was trained on the same database of 1000 SCA results as was the FFNN. For the CNN, the input is the micro-scale stress at each point (voxel) of the RVE, given by the cluster-wise results of SCA, σ(α,β) Each point in the RVE contains the three 2D stress components, like the RGB channels used for images. The output of the CNN is the macroscale strain εM that was applied as the loading conditions and caused the observed stresses. The training can be written as an optimization problem, as given in Eq. (1-31). The equations for training a CNN with padding, convolution, and pooling layers repeated Nconv times is given by:

find : W m n l , b n l ( l = 2 , 3 N L ) , in FFNN W ζ , ψ ζ , κ , η , b κ , n ( κ = 1 , 2 N kernel ) , ( η = 1 , 2 N conv ) , in CNN min loss function : MSE = 1 N T s = 1 N T ( ɛ M , s - ɛ * M , s ) 2 where : ɛ M , s = FFNN ( flatten ( pcp N c o n v ( p c p 2 ( p c p 1 ( σ s ( a , β ) ) ) ) ) ) ( 1 - 31 )

The term pcpηs(α,β)) is used to simplify the notation by combining the nested operations shown in FIG. 30. It is defined as


pcpηs(α,β))=poolingη(convη(paddingηs(α,β)))),

which includes terms for padding, convolution, and pooling. The remaining terms in the training problem are define as follows. The weight and bias in the FFNN are Wmnl and bml, and Wξ,ψζ,κ,η, bκ,η are the weights and biases in the convolution operations. The ground truth is ε+M,s and the estimate is εM,s. The number of training samples is defined as NT, and NN(l=NL) is the number of neurons in output layer. In this case, NN(l=NL) is three. The sampling is indexed by s.

The inputs to the CNN are the three stress arrays corresponding to the components of stress given by σs(α,β). The outputs are three strain values εjM,s(j=1,2,3). Just as for the FFNN, the mean squared error (MSE) is reduced gradually step by step using one of several optimization algorithms.

Prediction

Just as with the FFNN, we can define the function CNNs(α,β)) that describes the operation performed by the trained CNN, in this case


CNNs(α,β))=εM,s  (1-32)

In FIG. 31, a histogram for the error between predicted εM and the validation data set (again, the NT=150 samples generated above) is computed using Eq. (1-33). Most of the predictions made by CNN have an l2 norm less than 1×10−5, showing the CNN produce an accurate prediction of the strain state. The l2 norm illustrate the CNN network is able to make a proper prediction of the validation data. In FIG. 32, the correlations between the CNN prediction of applied strain and the reference solution of the three strain components is provided, using the same validation data sets as mentioned before. The solid black lines are the ground truth: all perfect predictions should lay on those lines. All three cases have correlation coefficients higher than 0.99, suggesting the trained CNN can provide a good accuracy in predicting applied strain. These show that the CNN can effectively map the stress contour to the applied external strain. Such a map may play an important role in linking microstructure information with macroscale information, e.g., connecting microstructure failure strength to a macroscale strain state. This will assist the inverse design problem where the optimum loading state is inferred using local information.


DifferenceCNNs=∥CNNs(α,β))−εM,s(SCA)2.,s=1,2, . . . ,NV  (1-33)

Convolutional Neural Network for Classification with Database Generated by SCA

The application of a CNN to material microstructure predictions is not limited to the sample problem shown above. Another example use of a CNN is as a microstructure classifier; this is similar to its common application in image classification. By using a microstructure mesh and an applied strain on the microstructure as the input, a CNN can be trained to predict whether the microstructure will become damaged.

Training

The training procedure for the classification CNN is given by

find : W mn l , b n l ( l = 2 , 3 N L ) , in FFNN W ξ , ψ ζ , κ , η , b κ , η ( κ = 1 , 2 N kernel ) , ( η = 1 , 2 N comv ) , in CNN min loss function : cross entropy = 1 N T s = 1 N T ( - ( d * s log ( d s ) + ( 1 - d * s ) log ( 1 - d s ) ) ) where : d s = FFNN ( flatten ( pcp conv ( pcp 2 ( pcp 1 ( σ s ( α , β ) ) ) ) ) ) nested operation : pcp 2 ( σ s ( α , β ) ) = 𝒫 pooling η ( 𝒞 conv η ( 𝒫 padding η ( σ s ( α , β ) ) ) ) ( 1 - 34 )

The output of the classification CNN, ds, is a binary indicator: 0 for non-damaged, 1 for damaged. Since the output is a binary value, the objective function is now defined in terms of the cross entropy between the truth value d*s and the predicted value ds. Cross entropy, or log loss, is widely used to measure the performance of a classification model. The CNN in this section is trained on the data extracted from the database used for the previous NNs. The database for training the classifier includes pairs of micro stress distributions and damage indicators. In this case, a critical-stress-based damage criterion is used to decide whether an RVE is damaged or not: if any von Mises stress σ(X) exceeds a critical von Mises stress σ* the RVE is considered damaged.

Prediction

Once trained, the CNN can predict whether the applied loading will result in microstructure damage without carrying out the full RVE simulation:

CNN classify ( σ s ( α , β ) ) = d s = ( 0 , Non - damaged 1 , Damaged ( 1 - 35 )

Such a CNN will be used in a microstructure-based topology optimization example to illustrate the effect of a microstructural damage constraint on the optimized structure.

Microstructure-Based, Multiscale Topology Optimization Using Neural Networks

In this section, we will illustrate how FFNNs and CNNs might be used in topology optimization to achieve microstructure-based design. This differentiates the current approach from classical topology optimization which typically uses simple constitutive relationships. As explained above, we propose to compress the RVE response database into an FFNN for forward prediction of RVE stress responses, where it will act similarly to a traditional homogenized constitutive model. However, because a database of RVE responses is used, no functional form of the homogenization is required. The RVE microstructure damage responses is represented with a trained FFNN+CNN; this introduces microstructure damage, linked with the applied strain state of the RVE. By using well-trained FFNNs and CNNs, two different optimization problems are defined as below:

1. Topology optimization with a material constitutive law extracted from an FFNN trained on the stress-strain relationship of a given microstructure. In this case, a non-linear material behavior during topology optimization is used to achieve a design that is durable under extreme loading conditions where the material response enters the non-linear region.

2. Topology optimization with constraints defined by the FFNN+CNN framework to identify microstructure damage and thereby design durable (damage aware) structures, using the CNN. In this case, the microstructure damage acts as an extra constraint to the topology optimization formulation to achieve a design that alleviates or avoids possible local microstructure damage.

In these two example problems, the design zone is described with a 60×30 mesh of rectangular, linear elements. Each element is a 1 cm×1 cm square. The elastic material properties are from the homogenized SCA results: E=200 MPa and v=0.27. An approximated non-linear elastic material response is extracted from the RVE simulation obtained from Sect. 3, as described in the following sections. Note that while plasticity is used, our FFNN is only valid for a non-linear elastic approximation of the material response. A point load of 75 N is applied at right bottom corner. The desired volume fraction of the optimized part is set as 0.35 of the original design zone. For the second case with damage the critical stress is defined as 0.7 MPa.

Topology Optimization with FFNN

The formulation of microstructure sensitive topology optimization with an FFNN is shown in Eq. (1-36). The objective function is defined as the overall strain energy of the structure, Φ, which is to be minimized. A subtle but important difference in the present example in this work is that the FFNN is used to replace the usual linear elastic material response with a nonlinear one. By using a nonlinear material responses database depicted in above, a new avenue for data-driven material and structure design is illustrated. This is depicted graphically in FIG. 33, which shows the use of an FFNN to generate microstructure-based stress-strain response within a topology design framework.

minimize : Φ = Ω M fu d Ω M + Ω M tu dS M , u U * with : Φ Ω M σ ( X M ) ɛ ( X M ) d Ω M , X M Ω M subject to : V ( ρ ( X M ) ) V * 0 ρ ( X M ) 1 , X M Ω M σ ( X M ) = ρ ( X M ) FFNN ( ɛ ( X M ) ) , X M Ω M FFNN for microstructural response , where : V ( ρ ( X M ) ) = 1 Ω M Ω M ρ ( X M ) d X M Ω M : macro domain , Ω : micro domain X M : coordinate in macro domain , X : coordinate in micro domain ( 1 - 36 )

where f is the body force, t is the applied traction on the boundary of the design zone, and u is the local displacement in the design zone. U* is the admissible displacement field, and SM defines the boundaries of the macrostructure (the design region). σ(XM) and σ(XM) are macro stress and strain. The density for each macro mesh is ρ(XM). The desired volume of material remaining in the design zone is V* and defined as 0.35, and V(ρ(XM)) is the optimized volume in the design zone. The density of location XM is ρ(XM) in the design zone, and the homogenized stress is σ(XM), defined as the product of ρ(XM) and FFNN(ε(XM). Here, FFNN represents a trained FFNN that will generate stress predictions based on given strain input ε(XM), as defined in Eq. (1-17)

In this case, the FFNN is used to approximate the RVE responses. Hence, σM is directly approximated by the FFNN. In truth, constraints of the software used for optimization require a functional description of the behavior (this limitation will be addressed in future work), thus the effective von Mises RVE stress versus effective strain curve is approximated using an exponential function: σM=0.7784*e22.18*εM−0.8071*e−378*εM. This hyperelastic material definition was fit to the FFNN results. This ad-hoc approach ensures stability of the optimization process by minimizing the overall strain energy using condition-based optimization.

FIG. 33 is topology optimization setup with FFNN. The FFNN is used to compute non-linear material responses to drive for a new design. This replaces the constitutive law commonly used for the macroscale with a homogenized response of the microstructure for each point in the macroscale. Mathematically, σM(XM)=ρ(XM)FFNN(ε(XM)),∀XMM, as defined in Eq. (1-36).

The optimized beam structures are illustrated in panels (a)-(b) of FIG. 34 for linear elastic material and non-linear elastic material, respectively. Necessary results are provided in Table 1-7. Two structures show substantial difference in the final shape of the structure. This means the material non-linearity plays an important role in topology optimization, where a new truss structure is realized in order to ensure minimal strain energy. The result suggests the importance of considerable of microstructure-based material non-linearity into structure optimization. It may also possible to include microstructure variation, such as different particle volume fractions, into the structure. In short, FFNN provides an alternative for a data-driven microstructure-based topology optimization, where the microstructural effect can be incorporated into the process and to achieve different designs that meet the design criteria.

TABLE 1-7 Results of the FFNN-based non-linear-elastic optimization problem. Note that substantial speedup is achieved while retaining the accuracy and microstructural basis of the concurrent approaches. This would provide further speed advantages in 3D. linear FFNN FE-SCA FE-FE material (nonlinear) concurrent* concurrent* Initial compliance 12.6 (6.3) 20.0 (10.0) (strain energy) (N cm) Optimized compliance 28.0 (14.0) 38.0 (19.0) (strain energy) (N cm) Database generation + 0 313 training (s) Optimization 338  472 23 328 220 × 106 calculation time (s) Factor of speed-up 280 255     9 431 over FE-FE No. of iterations 14   18 *estimated, assuming the same number of iterations as FFNN

Topology Optimization with Constraints Defined by FFNN+CNN

Topology optimization results may have high stress concentration zones. If not treated properly, the concentration zones may cause unexpected damage and affect the function of the structure. A traditional way to address this is to add stress or strain constraints to optimization. Previous authors have focused on optimization with stress concentration and singularities. In these previous studies, most damage criteria are only related to the macroscopic material model and do not consider micro structure mechanical behavior. In this work, an FFNN+CNN framework trained by the database generated with SCA provides a microstructure-based prediction of damage for the damage criterion.

In order to do this, the FFNN defined in Eq. (1-25) that performs the operation σ(X)=FFNNmicroM) is used. As mentioned above, this predicts local, rather than homogenized, stresses in the RVE. These local stress distributions serve as input to the CNN, which indicates whether or not damage has occurred. The optimization formulations are thus:

minimize : Φ = Ω M fu d Ω M + Ω M tu dS M , u U * U T KU = e = 1 N ( ρ ( X M ) ) p u e T k 0 u e subject to : V ( ρ ( X M ) ) V * 0 ρ ( X M ) 1 X M Ω M X M Ω M ( X M ) , X Ω ( X ) : CNN classify ( σ ( X ) ) = 0 Microscale damage criterion , where : σ ( X ) = FFNN micro ( ɛ M ) Microscale stress prediction , Ω M : macro domain , Ω : micro domain X M : coordinate in macro domain , X : coordinate in micro domain ( 1 - 37 )

where Φ=Σε=1n(ρ(XM))pueTk0ue is the compliance of the overall structure, N is the number of elements, p is the penalization power (typically p=3), ue is the displacement for each element, and k0 is the element stiffness. The averaged strain response of the RVE is εM. The local stresses within the RVE are defined as σ, and are calculated with a trained FFNN FFNNmicro. Other variables are the same as defined above for the FFNN-only optimization. For any X in RVE Ωm(X), the output value of CNNclassify should be 0, which represents a non-damaged state.

Since the FFNN+CNN database only gives a criterion, sensitivity analysis is not preferred in this case. The algorithm here follows a refined optimally criteria (OC) method. The optimization program structure is based on the 99-line topology optimization code (and thus the problem is similar, though the implementation is not identical to the FFNN-based optimization above). During the density update, for each element, three strain components will be passed to the FFNN+CNN. The FFNN+CNN will determine whether the current strain state is acceptable by assessing the local microstructural response. This is shown schematically in FIG. 35. For each material point within the design zone, the FFNN is used to compute the material response, be it linear or non-linear, considering the effect of microstructure. The CNN is used to incorporate microstructure damage, which will drive the optimization algorithm for a new design compared to a topology optimization with only linear material. Mathematically, this is ∀XM M (XM): σ(X)=FFNNmicroM(XM)), ∀X∈(X): d(XM)=(CNNclassify(σ(X))), as defined in Eq. (1-37). If any d(X) is marked as damaged in the microstructure, the XM point in the design zone containing that microstructure will be marked as damaged. If an element has been damaged, the density of this element will be increased by applying a penalty factor, while the densities of the rest of the elements will be decreased to satisfy the volume constraint.

Similarly to the FFNN example, panel (a) of FIG. 36 shows the reference case: a linear elastic optimization without a damage constraint. The final compliance is 30 N·cm. Panel (b) of FIG. 36 shows the optimized design with a microstructure-based damage constraint. The final compliance is 31 N·cm. Notice that while the compliance is quite similar, the design under a microstructural damage constraint resembles a more conventional truss structure; the optimization has avoided sharp angles and has fewer beams that give rise to stress concentrations likely to result in microstructure-driven damage.

A summary of the design variables and important parameters related to the optimization is provided in Table 1-8. The simple examples above show the potential application of an FFNN+CNN database generated by clustering reduce order methods. However, the optimization is just based on an artificially defined optimality criterion. The algorithm may not be stable for all kinds of problems. In the future, we should study the sensitivity and singularity of constraints based on an FFNN+CNN database.

TABLE 1-8 Results of the FFNN + CNN constraint optimization problem. Note that substantial speedup is achieved while retaining the accuracy and microstructural basis of the concurrent approaches. This would provide further speed advantages in 3D. Linear FE-SCA FE-FE material FFNN + CNN concurrent* concurrent* Initial compliance 295 (148) 295 (148) (strain energy) (N · cm) Optimized compliance 30.0 (15.0) 31.0 (15.5) (strain energy) (N · cm) Database generation + 0 512 training time(s) Optimization 12.6 14.5 69 674 660 × 106 calculation time (s) Factor of speed-up 45 × 106  9 473 over FE-FE No. of iterations 53 61 *estimated, assuming the same number of iterations as CNN

SUMMARY

Two challenges with current approaches to machine learning methods in the mechanical science of materials are: (1) the database generation time and effort are extensive, and (2) the application of machine learning is not well developed or understood by the community. This study covers several different topics related to these challenges:

    • We have outlined, related, and compared three different clustering-discretization methods (SCA, VCA, and FCA) that rely on unsupervised learning for order reduction and the solution of mechanistic governing equations for prediction.
    • One of these methods, SCA, was used to develop an example material behavior database suitable for training neural networks. This approach to database development substantially reduces the effort required to acquire the information upon which neural networks may be trained.
    • The basic operations, and how these combine to make predictions of mechanical responses, in an FFNN were outlined. This includes the role of weights, biases and activation functions as well as the description of the training stage of the neural network as a minimization problem using notation common within the mechanical sciences.
    • A similar description of convolutional neural networks was developed for two different possible applications: (1) to solve inverse problems where the boundary conditions need to be identified from a known stress distribution and (2) as a classifier to identify if damage will occur within a microstructure given a known stress distribution.
    • Two microstructure-sensitive topology optimizations are demonstrated. In the first case, the material response at the microscale derived from the FFNN results, and used to perform design against a load that causes the material to behave in a non-linear elastic way. In the second case, a material damage constrain is added to the optimization, where the CNN is used to identify if damage has occur on the microscale and penalize the design accordingly.

In short, we have provided methods to more rapidly produce the data needed to train neural networks, developed further insight into the working of neural networks from a mechanical sciences perspective, and highlighted the potential for these methods to enhance practical design tasks. The database of responses made with SCA, codes used for training and prediction with the neural networks, and the topology optimization codes are developed. This will encourage the use of data science and machine learning as a tool for mechanistic analysis, rather than simple as an unknown black-box operator.

Several areas where further investigation might be useful have already been noted:

    • Further development of clustering methods to represent large deformations, better capture anisotropic behavior or behavior that changes due to loading conditions, and even refine clusters during the prediction stage might be promising. The development of contact or self-contact formulations applicable to clustering discretization methods would aid in generality.
    • Formulations of clustering discretization solutions applicable to the component scale (rather that only the RVE scale), and the extension of concurrent multiscale solutions that use clustering discretization at multiple scales are currently under development.
    • For neural networks, methods to include history-dependence (e.g., plasticity) are currently an active area. Including physics in the neural network directly is another developing area, e.g., with physics-informed neural networks (PINNs).
    • For optimization, sensitivity analysis for topology optimization with FFNN and CNN should be further developed. More flexible software to support this would also be desirable.
    • Multiscale topology optimization with various material microstructure databases is still a developing area. The approach outlined here may be a promising method to simultaneously optimize topology and microstructure given sufficient constraints.
    • The “data-driven” component of these methods (both clustering-based discretization and neural networks) are not restricted to the use of computational data. Information from other sources, e.g., experimental sensor data and images, could be included if it is available. If mixed data streams are used extra care in data representation would be required.
    • Continuous validation and verification studies will help make these methods robust and reliable.

Example 2 Data Science for Finite Strain Mechanical Science of Ductile Materials

In this example, a mechanical science of materials, based on data science, is formulated to predict process-structure-property-performance relationships. Sampling techniques are used to build a training database, which is then compressed using unsupervised learning methods, and finally used to generate predictions by means of mechanistic equations. The method presented in this example relies on an a priori deterministic sampling of the solution space, a K-means clustering method, and a mechanistic Lippmann-Schwinger equation solved using a self-consistent scheme. This method is formulated in a finite strain setting in order to model the large plastic strains that develop during metal forming processes. An efficient implementation of an inclusion fragmentation model is introduced in order to model this micromechanism in a clustered discretization. With the addition of a fatigue strength prediction method also based on data science, process-structure-property-performance relationships can be predicted in the case of cold-drawn NiTi tubes.

Increasing research efforts in fine scale experiments and numerical modeling in recent decades have progressively led to a change in modeling approaches in mechanics and materials science. Empirical and phenomenological material laws that were previously used to model the nonlinear mechanical response of structures and materials are being replaced by microstructure-based mechanistic material laws. Under arbitrary loading conditions the number of microstructure observations and conditions to be modeled make the effort required for such an endeavor untenable for practical applications. The appeal of data science and in particular machine learning is a drastic reduction in the number of microstructure observations and simulations required to generate predictive material laws. There is hence a great interest in a data science theory for mechanical science of materials that could generate predictive material laws from a predefined database of experimental and numerical results.

Multiple approaches have been proposed in the literature to reach this goal, generally summarized by three steps: (1) collecting data using high-fidelity experiments and simulations to build a training database; (2) compressing the training database using unsupervised learning methods for dimension reduction; and (3) generating predictions using supervised learning methods or mechanistic equations on the compressed training database and optionally cross-validating those predictions using a testing database with new high-fidelity experiments and simulations.

The training database can be generated using, e.g., random sampling, Gaussian processes, or Sobol sequences. Because those sampling methods may require a lot of data points to cover the solution space sufficiently for accurate predictions, deterministic sampling methods have been considered by some authors. For instance, instead of considering a large number of arbitrary, random loading conditions for the training database, only 6 orthogonal loading conditions of small amplitude were proved to be sufficient for small strain elastoplastic analysis.

Compression of the training database can be achieved using various unsupervised learning methods for dimension reduction, such as Proper Orthogonal Decomposition (POD), K-means clustering and self-organizing maps. The choice of compression method has a significant importance as it defines the discretization of mechanistic equations that will be solved in the prediction stage. POD leads to shape functions of global support, while clustering methods ensure a cluster-wise discretization.

As a result of data compression, the complexity of high-fidelity experiments and simulations that were used to build the training database is encapsulated in a few degrees of freedom. In order to solve for those degrees of freedom and predict mechanical response at arbitrary loading conditions, mechanistic equations have to be reformulated in terms of the reduced degrees of freedom. This new formulation of mechanistic equations is usually called a reduced order model, although this denomination encompasses approaches such as proper generalized decomposition which do not rely on data science.

Additionally, some approaches couple the data compression and mechanistic prediction steps to improve the reduced order model during the simulation. Some supervised learning methods have been applied directly to the training database with a built-in compression stage. This is the case for instance for artificial neural networks, which have been applied in the literature to predict mechanical properties of materials as a function of their microstructural characteristics.

In this exemplary study, we disclose a data science mechanistic approach for ductile materials by Self-consistent Clustering Analysis (SCA), a data-driven mechanistic material modeling theory developed for small strain elastoplastic materials. SCA relies on data compression through clustering and mechanistic prediction through micromechanics and homogenization theory.

In this study, the mechanistic equations that SCA relies on to make predictions are reformulated for finite strain elastoplastic materials. Numerical convergence of this new method is verified. This new formulation of SCA enables the prediction of the nucleation of voids in ductile materials by debonding and fragmentation of inclusions at the scale of their microstructure, which is shown in FIG. 37, where ductile materials' microstructures are discretized using voxel meshes with matrix shown in blue and inclusions in red: panel (a) two-dimensional microstructure, and panel (b) inside view of a three-dimensional microstructure with a fragmented inclusion surrounded by a debonding void shown in light gray. This prediction is achieved with a complexity reduced by several orders. This advantage is exploited to predict process-structure-property relations for cold drawn Nickel-Titanium (NiTi) tubes.

Data Science Formulation

Microstructure-based material modeling requires the definition of an idealistic or statistically representative microstructure realization, called RVE. Homogenized material laws can be computed by analytically or numerically solving a boundary value problem for the response of that RVE. For arbitrary microstructure geometries and complex behavior of microstructure constituents (plasticity, fracture), numerical methods such as the Finite Element (FE) method or Fast Fourier Transform (FFT)-based numerical methods are required.

The microstructures that will be studied in the present paper correspond to ductile materials and feature one or multiple inclusions and voids embedded in a matrix, as shown in FIG. 37. The complexity of the microstructure's constituents' behavior arises due to the hyperelastoplastic response of the matrix, the hyperelastic-brittle behavior of the inclusions, and debonding micromechanisms at the matrix/inclusions interface.

The FE method can be used with any structured or unstructured FE mesh of the undeformed RVE domain Ω0m (the superscript m means microscopic), while FFT-based methods require structured voxel meshes such as that shown in FIG. 37. In the FE method discrete equations are written for the displacement field um, which is approximated at mesh nodes as


um(X)≈Σn=1Nnodesum,nNn(X),X∈Ω0m,  (2-1)

where Nnodes is the number of nodes in the FE mesh, um,n is the displacement vector at node n, and Nn is the FE shape function at node n. In FFT-based numerical methods, discrete equations are written for the deformation gradient tensor field Fm=I+∇Xum, which is approximated voxel-wise as


Fm(X)≈Σn=1NvoxelsFm,nχn(X),X∈Ω0m,  (2-2)

where Nvoxels is the number of voxels, Fm,n is the deformation gradient tensor in voxel n, and χn(X) is the characteristic function which is equal to 1 if X is inside voxel n and zero otherwise.

For a given microstructure, the displacement field um and the deformation gradient field Fm depend on boundary conditions applied to the RVE. In the present work, um will be decomposed over the RVE domain Ω0m into a linear part and a periodic part. As a result, Fm will be decomposed into a constant part FM (the superscript M means Macroscopic) and a periodic part with zero average over Ω0m. These assumptions correspond to first order homogenization theory.

Data science is used in the mechanical science of materials to predict either um or Fm as a function of FM. As stated in the introduction, the first step is to generate data through simulations. Simulation results in the training database will have large dimensions due to dependence of approximations in Eqs. (2-1) and (2-2) on either the number of nodes or the number of voxels. Data compression is necessary to obtain new approximations with reduced dimensions.

Data Compression

Dimension reduction can be achieved using various methods among which POD and clustering are presented and compared in the following.

The general formulation of data science approaches that is developed herein is only relevant if the complexity of simulations that are to be conducted in the prediction stage is at least one order superior to the complexity of simulations required in the training and data compression stage. The relevance of data science approaches also depends on the amount of work that can be transferred out of the prediction stage. This will be evidenced in the following in the case of POD and clustering based data science approaches for mechanical science of materials.

Data compression in the case of POD consists in replacing the large number of local FE shape functions (Nn)n=1 Nnodes by K<<Nnodes global functions (Wik)k=1 K,i=13, called principal components or modes. The latter can be computed using various decomposition techniques such as principal component analysis or singular value decomposition. The resulting approximation replacing Eq. (2-1) is


uim(X)≈Σk=1Kuim,kWik(X),X∈Ω0,  (2-3)

where the modes are discretized at mesh nodes as


Wik(X)=Σn=1NnodesWik,nNn(X),X∈Ω0.  (2-4)

Simulations in the prediction stage can then be conducted using a standard FE weak form but replacing approximation of Eqs. (2-1) by (2-3). It is interesting to see that once the modes are computed in the data compression stage, Eq. (2-4) can be precomputed at integration points of the FE mesh in the same way that FE shape functions are usually precomputed in FE codes.

However, if the material is heterogeneous, or if it has a nonlinear behavior that leads to heterogeneous deformations, material integration still has to be solved at each integration point. Consequently, in the POD method, the complexity of material integration is not reduced. Additionally, the stiffness matrix associated to the FE weak form is dense because of the form of Eq. (2-4), and hence its solution using direct or iterative solvers has a cubic worst-case complexity instead of quadratic. However, this complexity depends on K instead of Nnodes, with K>>Nnodes, and is therefore drastically reduced by POD.

Data compression in the case of SCA follows a different approach, where the initial numerical method is FFT-based. The large number of voxels is to be replaced by K<<Nvoxels mutually-exclusive groups of voxels that are called clusters and that span the entire RVE domain. Clusters can be constructed using various clustering techniques such as K-means clustering, or self-organizing maps. Examples of data that can be used for clustering are given below. The resulting approximation replacing Eq. (2-2) is


Fm(X)≈Σk=1KFm,kχk(X),X∈Ω0m,  (2-5)

where Fm,k is the cluster-wise constant deformation gradient tensor in cluster k, and χk(X) is the characteristic function which is equal to 1 if X is inside any voxel of cluster k, and zero otherwise. Because in the FFT-based numerical method the degrees of freedom are directly the voxel-wise constant deformation gradients, interpolation and integration are carried out at the same points. Thus, clustering degrees of freedom directly leads to a reduction of the number of degrees of freedom and of material integration complexity. In fact, in SCA, the complexity of all operations conducted in the prediction stage only depends on the number of clusters K, with the most expensive operation being, similarly to POD, the solution of a dense linear system. The latter results from the reformulation and discretization of the Cauchy equation into the discrete Lippmann-Schwinger equation. These steps are described in the following in the finite strain case following recent work on finite strain FFT-based numerical methods and then integrating it into SCA.

Continuous Lippmann-Schwinger Equation

As mentioned previously, first order homogenization consists in defining the deformation gradient tensor field in the RVE Fm as the addition of the macroscopic (homogeneous) deformation gradient FM and a microscopic (heterogeneous) fluctuation. Hill's lemma can be used to define the macroscopic first Piola-Kirchhoff stress tensor PM as the average of the microscopic one

P M = 1 Ω 0 m Ω 0 m P m ( X ) dX .

Hill's lemma requires (Fm−FM) to verify compatibility, i.e., to derive from a periodic displacement field, and Fm to verify equilibrium, i.e., to be the solution of the Cauchy equation


X·Pm(Fm(X))=0,X∈Ω0m.  (2-6)

It can be shown that Eq. (2-6) is equivalent to the Lippmann-Schwinger equation

F m ( X ) = - Ω 0 m 𝔾 0 ( X , X ) : ( P m ( F m ( X ) ) - 0 : F m ( X ) ) dX + F 0 , X Ω 0 m . ( 2 - 7 )

The fourth rank tensor 0 is the stiffness tensor associated to an isotropic linear elastic reference material. The far field deformation gradient tensor F0 and the periodic Green's operator 0 are determined below. The latter maps any tensor field τm to a compatible one:


u∈(H10m))3,u periodic on Ω0m,−0m=∇xu,  (2-8)

where H10m) is the Sobolev space of square-integrable functions whose weak derivatives are also square-integrable.

The combination of Eqs. (2-7) and (2-8) yields a microscopic deformation gradient tensor Fm that verifies compatibility and a first Piola-Kirchhoff stress tensor Pm that verifies equilibrium.

Discrete Lippmann-Schwinger Equation

SCA includes solving Eq. (2-7) cluster-wise instead of voxel-wise. This choice is inspired from micromechanics and in particular Transformation Field Analysis. FIG. 38 shows an example of clustering performed on the microstructures in FIG. 37. In FIG. 38, panel (a) shows two-dimensional microstructure discretized using 8 clusters; panel (c) shows same two-dimensional microstructure discretized using 65 clusters; and panel (c) shows three-dimensional microstructure discretized using 217 clusters showing two clusters in the matrix phase (two shades of blue), one cluster in the inclusion phase (red), and one cluster in the void phase (light gray).

As a result of the training stage, the RVE domain Ω0m is discretized into K subsets (Ω0m,k)k=1 K. The degrees of freedom in the FFT-based numerical method are associated with the microscopic deformation gradient Fm. In SCA, Fm is discretized by a cluster-wise constant approximation (Fm,k)k=1 K. As a consequence, the microscopic first Piola-Kirchhoff stress tensor is also approximated cluster-wise (Pm,k)k=1 K, and Eq. (2-7) can be discretized as


Fm,k=−Σk′=1 K0,k,k′:(Pm,k′0:Fm,k′)+F0,k=1K  (2-9)

where 0 is the interaction tensor defined by

𝔻 0 , k , k = 1 Ω 0 m , k Ω 0 m χ k ( X ) Ω 0 m χ k ( X ) 𝔾 0 ( X , X ) dX dX = 1 Ω 0 m , k Ω 0 m , k ( χ k * 𝔾 0 ) ( X ) dX . ( 2 - 10 )

The characteristic functions χk and χk′ are equal to 1 in, respectively, clusters k and k′, and 0 elsewhere. In the FFT-based numerical method, the periodic Green's operator 0 depends on 0, and is known in closed form in Fourier space. Because 0 is related to an isotropic linear elastic reference material, 0 can be expressed in Fourier space as a function of the reference Lamé parameters λ0 and μ0. It is then obtained in real space by using the inverse FFT. In particular, Eq. (2-10) can be written in the form

𝔻 0 , k , k = f 1 ( λ 0 , μ 0 ) 𝔻 1 , k , k + f 2 ( λ 0 , μ 0 ) 𝔻 2 , k , k , 𝔻 i , k , k = 1 Ω 0 m , k Ω 0 m , k FFT - 1 { FFT { χ k } 𝔾 ^ i } ( X ) dX , i = 1 , 2. ( 2 - 11 )

The detailed expressions of f1, f2, 1 and 2 can be found among others. Drastic computational cost reduction is enabled by SCA thanks to a reduced number of degrees of freedom by clustering, and by the fact that 1 and 2 can be precomputed in the training stage. Therefore, neither FFTs nor inverse FFTs are computed in the prediction stage, even if the reference material is changing.

In the present work, mixed boundary conditions are coupled to Eq. (2-9). Some components Fi,jm of the average of the microscopic deformation gradient are set equal to their macroscopic counterparts from Fi,jM, and some other components Pi,jm of the average of the microscopic first Piola-Kirchhoff stress tensor are set to zero. This can be done by adding the following conditions to Eq. (2-9):

{ k = 1 k Ω 0 m , k F i , j m , k = Ω 0 m F i , j M , ( i , j ) k = 1 k Ω 0 m , k P i , j m , k = 0 , ( i , j ) { 1 , 2 , 3 } 2 ( 2 - 12 )

where ⊂{1,2,3}2 is the set of components for which kinematic conditions are imposed.

As noted, solutions of Eq. (2-9) depend on the choice of reference material. An optimal choice can be computed in the prediction stage by making the reference material consistent with the homogenized material. This means that the far field deformation gradient tensor F0 is an additional unknown that must be solved for in SCA, as opposed to the FFT-based numerical method where F0≡FM. The self-consistent method consists in using a fixed-point iterative method where, at each step, the reference Lamé parameters λ0 and μ0 are changed so that ∥PM0: (F0−I)∥2 is minimized. This is presented in algorithm form in the Appendix.

To summarize, SCA is based on a voxel-wise discretization of the RVE domain, which is inherited from FFT-based numerical methods. The originality of SCA comes from the use of a K-means clustering algorithm in the training stage to cluster voxels based on a mechanistic a priori clustering criterion computed using a simple sampling of the loading space. This training stage also includes computing all voxel-wise and computationally expensive operations such as FFTs and inverse FFTs.

In the prediction stage, a self-consistent iterative algorithm is used to search for the optimal choice of reference Lamé parameters. At each iteration of this self-consistent loop, matrix assembly operations are accelerated because all voxel-wise operations have been precomputed in the training stage and already reduced to cluster-wise contributions. A Newton-Raphson iterative algorithm must be embedded within each self-consistent iteration as we are considering nonlinear materials, thus the discrete Lippmann-Schwinger equation is linearized.

The outputs from SCA are the microscopic variables' cluster-wise approximations, including the microscopic first Piola-Kirchhoff stress tensor. The latter can be used to predict void nucleation micromechanisms through stress-based fracture criteria as will be described in Sec. 4.

The main advantage of SCA over POD techniques is that in the prediction stage all operations are conducted cluster-wise in SCA instead of voxel-wise, including material integration and even fragmentation modeling.

Numerical Validation

Before considering a specific application, general ductile materials microstructure are modeled in this section both using a finite strain FFT-based numerical method and finite strain SCA. The goal is to validate the numerical convergence of SCA towards the reference result, and its capability to compute accurate predictions with a reduced complexity, as was shown for the small strain case.

In the present large strain case, hyperelastic behavior is modeled both in the matrix and in the inclusions. A multiplicative von Mises plasticity model with linear isotropic hardening is added to model the nonlinear response of the matrix. Inclusions are assumed to be brittle, which is a common assumption for hard phases in ductile materials. The model ductile material microstructure is shown in panel (a) of FIG. 39, and material properties are given in Table 2-1 for the matrix and the inclusions. The microstructure is discretized using 100×100×100 voxels, which is sufficient to accurately predict the response with the FFT numerical method based on our preliminary calculations (not reported herein).

TABLE 2-1 Material parameters for the model ductile material Parameter Matrix Inclusion Void Units Young's 70.0 400.0 70.0e−3 GPa Modulus Poisson's Ratio 0.33  0.2 0.0 Yield Strength 400 MPa Hardening 1333 MPa Modulus

The reference result is computed on the voxel mesh using the finite strain FFT based numerical method under unidirectional tension up to 25% applied logarithmic strain with strain increments of size 0.001. The training database used for K-means clustering simply includes the voxel-wise deformation gradient tensor extracted from the first increment of this reference simulation, which corresponds to a linear elastic analysis. In other words, the simulation used for training is identical in all aspects (mesh, geometry, material properties) to the one conducted in the prediction stage, except for loading conditions because only one strain increment is applied during training. This database can hence be constructed with a negligible computational cost. The K-means algorithm requires a predefined number of clusters, which is varied from k1=1,4,16,64,256 in the matrix phase, and k2=1,1,4,13,26 in the inclusion phase.

The comparison between the reference macroscopic stress/strain curve and those predicted by finite strain SCA is presented in panels (b)-(c) of FIG. 39. For 16 clusters in the matrix and more, a very accurate prediction is obtained with finite strain SCA. This extends the validation conducted to large strains.

To show the influence of clustering on computational complexity, a comparison of computation times is presented in panel (c) of FIG. 39. The computation time is shown to be reduced by four orders using SCA compared to the FFT based numerical method. This shows that SCA drastically reduces the complexity of microstructure calculations, based on a mechanistic clustering of voxels.

This interesting advantage of SCA is demonstrated in a second set of simulations where the inclusions in panel (a) of FIG. 39 are replaced by voids. Material properties for this porous ductile material are the same as those in Table 1, except that the stiffness of the voids is assumed to be 1% of that of the matrix. Loading is set to uniaxial tension to ensure that the stress state remains constant during the analysis.

The comparison between the reference macroscopic stress/strain curve and those predicted by finite strain SCA is presented in panels (a)-(b) of FIG. 40. It can be seen that convergence is much slower for this example with voids, which features larger plastic strains than in the inclusions case. However, predictions are very close to the reference result.

The main advantage of SCA in terms of computation time is preserved, as shown in panel (b) of FIG. 40. As all SCA predictions are very close, one could use a very small number of clusters and obtain a good approximation of the reference result, resulting once again in a drastic reduction of computational complexity.

These first simulations with the proposed finite strain SCA formulation are promising as predictions are close to the reference results for a drastically reduced computational complexity. However, this comparison is purely global: only averaged results are being assessed. At large plastic strain or for more complex loading conditions, very heterogeneous and localized strain fields may develop and would require improvements to the method. For instance, adaptive clustering techniques could be considered to update the clustering when plastic localization phenomena occur within the RVE.

Fatigue Strength Prediction of Cold Drawn NiTi Tubes

The objective of this section is to demonstrate the utility of a mechanical science of materials based on data science to predict process-structure-property-performance relationships. The chosen application is the prediction of the fatigue strength of cold drawn NiTi tubes as a function of drawing ratio and initial inclusion Aspect Ratio (AR).

NiTi tubes are formed using a series of hot and cold metal forming processes coupled to heat treatments. They are used in the making of, e.g., arterial stents and heart valve frames that undergo a large number of cyclic loads due to heart beats. A critical measure of a tube's performance is hence the fatigue strength of its material, which is itself a function of the material's fatigue life under different applied cyclic strain amplitudes. This fatigue life can be predicted using micromechanical simulations, which depend on the microstructural constitution of NiTi tubes. The latter is a consequence of forming processes, and in particular of the cold drawing step, which is the focus of the following study.

We disclose simulating cold drawing at the microscale by applying biaxial compression to an initially debonded inclusion embedded within a NiTi matrix. In order to predict the evolution of this microstructure during cold drawing, the finite strain SCA theory introduced above is completed with a fragmentation model. The microstructure is extracted at different stages of this drawing model and used as input to a data-driven fatigue life prediction model developed in previous studies and described briefly. The transfer of the microstructure morphology from the drawing model to the fatigue life prediction model requires a displacement reconstruction and microstructure interpolation step that is described. Process-structure-property-performance predictions obtained using this data science approach are presented.

Cold Drawing Model

Using the self-consistent scheme presented above, the discrete Lippmann-Schwinger equation (2-9) can be solved with mixed boundary conditions (2-12) and appropriate constitutive models for the microstructure's constituents. Constitutive models and material parameters are kept identical to those used and reported in Table 2-1, but they are completed by an inclusion fragmentation model. In addition, inclusions are assumed to be initially debonded as a result of high shear stresses developing early at the inclusion/matrix interface during cold drawing, similarly to the cold extrusion process.

Inclusions fragmentation is modeled using a Tresca yield criterion averaged inclusion-wise, following a regularization technique used in a previous work and coupled to a size-effect criterion. The Tresca criterion defines the shear stress σTresca as


σTresca=1/2 max(|σ1−σ2|,|σ2−σ3|,|σ3−σ1)  (2-13)

where σ1, σ2, σ3 are the cluster-wise constant principal stresses computed within each cluster of the inclusion phase. The shear stress σTresca is then averaged inclusion-wise, or inclusion fragment-wise if the inclusion has already broken-up, and that averaged value σTresca is compared to the inclusion shear strength σTrescac. The equivalent radius of this inclusion or inclusion fragment r is also compared to a critical size parameter rc. If the shear strength has been reached and r≥rc, the inclusion cluster containing the σTresca-weighted barycenter of that inclusion or inclusion fragment is turned into void, as illustrated in FIG. 41. In practice, this consists in reducing its Young modulus to a 1000th of its initial value over several load increments. This procedure is carried out at the end of each load increment of the finite strain SCA simulation, once Eqs. (2-9) and (2-12) have been solved using the self-consistent scheme. As revealed in FIG. 41, the orientation of the fragmentation crack is predefined and is initially orthogonal to the drawing direction. The fracture criterion parameters are given in Table 2-2.

TABLE 2-2 Fracture criterion parameters for the cold drawing model Parameter Value Units Shear Strength 3000 MPa Critical 0.17 1/RVE size inclusion size

Fatigue Life Prediction Model

High-cycle fatigue life can be estimated based on local plastic deformation predicted under cyclic loading, where these deformations are computed using the SCA outlined above. Since cyclic strain amplitudes are low, typically below 1% reversed strain, a small strain formulation can be used. An appropriate micro-scale material law-crystal plasticity (CP)—is solved cluster-wise to obtain the cyclic change in plastic shear strain (Δγp) and stress normal (σn) to that strain in the matrix material. The peak value of these variables reach cyclic steady-state rapidly, typically within 3 or 4 loading cycles. From these, a scalar value called the Fatigue Indicating Parameter (FIP) can be defined, which quantifies the fatigue driving force at any location. Here, we adopt the Fatemi-Socie FIP, defined by

FIP = Δ γ p 2 ( 1 + κ σ 𝔫 σ y ) ( 2 - 14 )

This FIP is a critical plane approach based upon the plane of maximum normal stress, σn, normalized by the material yield stress (σy) and a material dependent parameter κ, which controls the influence of normal stress (here κ will be taken as 0.55). When used with SCA and a CP law, the FIP in each cluster is computed cluster-wise from plastic strain and stress across time increments using a simple search to maximize the plastic strain, and thus identify the critical plane. The maximum, saturated FIP (NFIPmax) can be correlated to the number of fatigue incubation cycles (Nine) a microstructure can withstand using a Coffin-Manson parameterization. By computing Ninc for a number of different strain amplitudes to generate a strain-life curve, the fatigue strength (strain amplitude at which a given number of cycles can be reached) can be computed through interpolation or fitting.

The CP material law, calibrated to capture the hardening response of the B2 phase of NiTi, is used. Post-yield hardening is computed with a backstress term that accounts for direct and dynamic hardening. Consistent with the worst-case or nearly worst-case approximation for the material phase, and the lack of crystallographic information from the cold drawing model, the matrix is assumed to be composed of a single grain oriented such that its Schmid factor is maximized. The hard inclusion phase (representing oxides or carbides) is taken to be linear-elastic with an elastic modulus 10 times that of the matrix material. This process follows the CPSCA formulation integrates the CP material law within SCA, and the fatigue prediction method shown for synthetic microstructures. Indeed, the same model parameters are used for CP, FIP and Coffin-Manson as are shown.

Transfer of Microstructure from Cold Drawing Model to Fatigue Life Prediction Model

The fatigue life prediction model relies on CPSCA and hence on an underlying voxel mesh. In order to use this model, the deformed and fragmented microstructures from the drawing simulation results need to be transferred to a new voxel mesh.

The first step is to reconstruct the microscopic displacement vector field, since both FFT-based numerical methods and SCA only solve for the microscopic deformation gradient tensor. This is done using a simple Taylor expansion, or forward finite difference formula, which computes the displacements at all nodes of the voxel mesh from the microscopic deformation gradients inside voxels. This computation starts at some arbitrary corner of the RVE domain where the displacement is fixed to zero, which is in agreement with FE-based linear homogenization implementations.

Once the displacement vector field has been reconstructed, the mesh can be deformed by adding these displacements to node coordinates, as shown in panels (a)-(b) of FIG. 42. Finally, the phase tags (matrix, inclusion, void) are transferred from this deformed mesh to a new voxel mesh of the deformed RVE domain using a simple voxel-wise constant interpolation. As a result, the new voxel mesh is compatible with FFT-based numerical methods and SCA, but embeds phase tags that correspond to a cold drawn microstructure. Fatigue life predictions can be computed on this new voxel mesh using the CPSCA method, as shown in panel (c) of FIG. 42.

Fatigue Strength Predictions

Three different initial conditions for the drawing model were considered, representing possible variability in the quality and degree of processing in the feed stock used for cold drawing. Each condition includes a single, ellipsoidal, debonded inclusion centered in the matrix with a different AR in the load direction. The cross-sectional area (i.e., the minor axis of the ellipsoid) of the initial configuration is kept constant, and the length (in the drawing direction) changed. By doing this, we study the influence of AR (mean curvature) on drawing, fragmentation and subsequent fatigue life. These three different cases are deformed to up to 60% section height reduction by applying biaxial compression with a stress-free third axis. For each case, at every 5% height reduction the procedure outline is performed and the fatigue strength is computed. In order to compute the fatigue strength, fatigue lives at strain amplitudes ranging between 0.36% and 0.54% in increments of 0.06% were computed, and the strain amplitude required to reach 107 cycles was estimated using a power-law fit to that data. Triangle waveforms with loading rate 0.1/s were used throughout to apply fully reversed tension-compression (strain ratio, R=−1) fatigue loading.

The results from this parametric study are shown in FIG. 43. In the center of the figure, fatigue strength—the maximum strain amplitude at which a specified number of cycles can be obtained—is plotted versus the section reduction resulting from drawing. At each point in section reduction and for each AR the fatigue strength at 107 cycles is computed using the procedure outlined above. Cross-sections of the volume are given at six different points in order to visualize and analyze the process by which the fatigue strength changes with reduction percent. The drawing model captures the overall trend of an increase in fatigue life with increasing height reduction, particularly for the highest AR. A large increase in the fatigue life is noted when the AR 3 particle fragments, between 40% and 45% reduction. The pre-fragmentation configuration is shown in the upper right-hand subset to FIGS. 2-7, and the post-fragmentation configuration is shown in panel (c) of FIG. 42. The distribution of FIP changes noticeably between these two states, with the field much more concentrated at the interfaces before fragmentation and more distributed throughout the volume after fragmentation. Up to 60% height reduction neither AR 1 nor 2 fragment, and no substantial decrease in life is noted. This is consistent with experimental experience, where relatively large reductions are often required to achieve large gains in fatigue performance. Higher reduction percents or different material properties would be required to fragment these cases.

According to the embodiment, a general formulation of data science approaches for mechanical science of materials is presented. This general formulation includes reducing the complexity of process-structure-property-performance prediction methods using unsupervised learning methods on a training database of high-fidelity simulations. This is evidenced in the case of a training database composed of RVE simulations results computed under various loading conditions using the FE method or FFT-based numerical methods. Dimension reduction leads to a compressed RVE model where nodes and voxels are replaced by either modes or clusters depending on the supervised learning method used for data compression.

In the prediction stage, supervised learning methods or mechanistic equations are solved using the compressed RVE. For instance, POD can be used to solve the Cauchy equation using a compressed FE discretization where the complexity depends on the number of modes instead of the number of nodes. Similarly, K-means clustering can be used to solve the Lippmann-Schwinger equation with a complexity depending on the number of clusters instead of the number of voxels. The interesting advantage of this later approach over the former is that it reduces both the complexity of mechanistic equations solution and material integration.

The solution of the Lippmann-Schwinger equation using a clustered discretization requires a self-consistent scheme that has been extended to finite strain elastoplastic materials and coupled to micromechanical void nucleation models in this example. As a result, microstructure evolutions due to large plastic strains have been captured during the cold drawing of NiTi tubes. These microstructure evolutions have been related to the fatigue life and then the fatigue strength of NiTi tubes using a second data science based approach developed in a previous work. Therefore, it has been demonstrated that the proposed data science formulation for mechanical science of materials can predict process-structure-property-performance with a reduced complexity.

Example 3 Predictive Multiscale Modeling for Unidirectional Carbon Fiber Reinforced Polymers

This exemplary study presents a predictive multiscale modeling scheme for Unidirectional (UD) Carbon Fiber Reinforced Polymers (CFRP). A bottom-up modeling procedure is discussed for predicting the performance of UD structures. UD material responses are computed from high fidelity Representative Volume Elements (RVEs). A data-driven Reduced Order Modeling (ROM) approach compresses RVEs into Reduced Order Models so material responses can be computed in a concurrent fashion along with the structural level simulation. The approach presented in this example is validated against experimental data and is believed to provide design guidance for future fiber reinforced polymers development process.

In modern engineering applications, composite materials are receiving growing attention for their extraordinary lightweight and strength. To understand the mechanical performance of various CFRP designs, physical tests are necessary. With computational power continually growing, it is now possible to utilize the Integrated Computational Material Engineering (ICME) approach to virtually evaluate the performance of composite designs and provide design guidance for composite materials. The ICME approach directly integrates microstructure information into property and performance prediction. In the ICME process, the intrinsic relationship between material microstructure and mechanical performance can be captured by a multiscale model which links microstructure to macroscale performance. In this exemplary study, a bottom-up ICME modeling framework for UD CFRP is introduced. The framework incorporates a two-stage ROM technique that enables rapid computation of UD microstructure nonlinear responses during UD CFRP structures simulation. The bottom-up multiscale modeling workflow for UD CFRP is explained in FIG. 44, which illustrates the length scale span in the modeling process. The UD CFRP microstructure at micron scale, characterization by UD RVE, provides microstructure information for UD lamina and UD coupon at millimeter scale. Accurate UD structure responses are predicted using physical RVEs. From the microscale (the UD microstructure) to the macroscale (the UD Coupon FE mesh), the UD Coupon is modeled in a bottom up fashion. This framework can be extended to 3-scale composites, including UD, woven and woven laminate structures.

There has been considerable effort during the past decades in incorporating microstructure information to the macroscale model for performance prediction. For example, one can model all microstructure details into a single model, but this computation is expensive due to the fine mesh required. By deploying multiscale modeling techniques, the macroscale responses are predicted with physical microscale information. A hierarchical multiscale modeling approach as demonstrated by the Multiresolution Continuum Theory is one of them. In the Multiresolution Continuum Theory, the microstructure information is implemented into the macroscale constitutive law to construct a hierarchical multiscale model, which captures microstructural effects, such as the effect of inclusion size. Although this hierarchical multiscale modeling method preserves certain microstructure information, it does not provide explicit microstructure evolution.

To capture microstructure responses, concurrent homogenization can be deployed. One of the concurrent homogenization schemes is the Finite Element square (FE2) approach. In the FE2 approach, the macroscale geometry is discretized with a Finite Element (FE) mesh. Material responses at all integration points are computed by solving RVEs that are discretized with FE meshes. The FE2 approach solves two sets of FE meshes in a concurrent fashion, which results in high computational cost.

To improve computational efficiency, various ROM approaches were developed, such as Transformation Field Analysis, Nonuniform Transformation Field Analysis and Proper Orthogonal Decomposition. A newly proposed data-driven two-stage ROM approach, namely Self-consistent Clustering Analysis (SCA), creates ROMs from high fidelity voxel mesh RVEs and simulates RVEs' elasto-plastic behaviors with an effective RVE damage and failure. In the offline stage, a three-step approach is introduced: 1) data collection, such as collecting strain concentration tensor for each voxel in the RVE; 2) Unsupervised learning, which classifies all voxel elements into different clusters; 3) Generating cluster-wise interaction tensors. In the online prediction stage, cluster-wise strain responses are identified by solving a discretized Lippmann-Schwinger equation. A previous publication shows that the SCA drastically reduces computational expenses, and the accuracy is verified against a high fidelity Direct Numerical Simulation (DNS). Therefore, the three-step two-stage data-driven SCA method is a valuable tool in the ICME process for modeling UD CFRP composites. An SCA flow chart for the solving the Lippmann-Schwinger equation in one embodiment is shown in FIG. 66.

In this exemplary example, the ICME modeling framework for UD CFRP structure performance prediction is presented. Under the framework, the macroscale UD model is discretized with an FE mesh. The UD microstructures are characterized by RVEs. RVEs are compressed into UD Reduced Order Models (ROMs) and provide mechanical responses for all integration points on-the-fly. The constitutive response of each constituent, e.g., fiber and matrix, is obtained from physical tests. The UD ROMs capture elastic and elasto-plastic responses of the UD CFRP through computing the RVE responses. Moreover, the UD non-linearity due to matrix plasticity differentiates this work from previous efforts in CFRP ICME modeling. In those previous works, structural analysis is made by assuming linear elastic material responses. The UD ROMs compute RVE responses efficiently enough to replace the phenomenological anisotropic material model and minimized material constants calibration effort.

The methodology developed can be popularized for predicting performance of fiber reinforced polymer in general. Basic materials information, the experimental procedures of the UD CFRP coupon off-axial tensile test and the UD CFRP 3-pt bending test, ICME modeling process for the UD CFRP in details, and results and experimental validation are discussed below.

Materials Information

In this section, the material properties obtained for A42 carbon fiber from DowAksa and thermoset epoxy resin from Dow chemical are provided. Material properties provided in this section are used through out the modeling work in this exemplary study.

The fibers elastic constants are given in Table 3-1 below. Fibers are assumed to behave elastically. The fiber direction compressive strength is assumed to be one-fourth of the tensile strength, where the tensile strength (TS) and compressive strength (CS) are assumed to be 4200 MPa and 1050 MPa, respectively. Ductile damage model for carbon fibers is assumed, as shown in Eq. (3-1), where dfiber=0 means no damage and dfiber=1 means fiber is damaged.

TABLE 3-1 Carbon Fiber Elastic Material Constants E11 E22 = E33 G12 = G13 GPa 19.8 GPa 29.2 GPa G23 v12 = v13 v23 .92 GPa 0.28 0.32

d fiber = ( 0 , σ 11 TS / CS 1 , σ 11 > TS / CS ( 3 - 1 )

The epoxy resin elastic constants and tensile and compressive strengths are given in Table 3-2. The tensile and compressive yielding curves are given in panels (a)-(b) of FIG. 45, respectively. Different tensile and compressive behavior suggest that a paraboloid yielding criterion can be implemented to capture such behavior.

TABLE 3-2 Material Constants of Epoxy Matrix E v σT σC 3803 MPa 0.39 68 MPa 330 MPa

Experiments for the UD CFRP

Material characterizations provide a good understanding of the material of interest. Various testing methods deliver the CFRP properties and provide validation data for the prediction made by the ICME framework. In this work, two types of experiments are identified for examining the predictivity of the proposed UD CFRP ICME framework: (1) UD coupon 10° off-axial tensile test and (2) UD 3-pt bending test.

UD CFRP Coupon 10° 10 Degree Off-Axial Tensile Test:

The UD CFRP coupon specimen is prepared through the following steps. The edge of the UD plaque (with nominal fiber volume fraction of 50%) served as reference for the determination of the angle for cutting the 10° off-axis orientation. The specimen head areas and the tab (woven fiberglass in an epoxy resin) surfaces (with a length of 50 mm) were prepared with grinding paper before applying a commercially available acrylic adhesive. Metallic wires with a diameter of approximately 220 m were used as spacers between specimen and the tab surfaces. The tabbing angle of about 16° was formed by grinding. The specimens were cut with a waterjet system to a nominal width (w) of 12.7 mm with a length of 210 mm for a resulting aspect ratio of 9 in the gauge section of length 1=120 mm, as shown in FIG. 46 (specimens 001-005). An abrasive grit size of 220 m and the lowest translation speed was chosen for minimal fabrication damage, based on preliminary studies for optimization of the waterjet cutting parameters (i.e., nozzle diameter, pressure, speed, abrasive grit size, etc.).

The displacement-controlled tensile tests have been conducted on a servo hydraulic testing machines at a quasi-static loading rate of 0.0167 mm/s (1 mm/min). The loading has been induced by the actuator located at the bottom of the test frames. Before each test, a precision steel block had been used for rotational alignment of the actuator to reduce out-of-plane misorientation. The specimens were rigidly hydraulically gripped with anti-rotation collars installed using diamond jaw surfaces and a pressure of 4 MPa. The gripping length on each side ranged between 30 mm to 40 mm. The specimens have been aligned with specimen stops in the grip.

All specimens were prepared for Digital Image Correlation (DIC) measurement with commercially available matte white spray paint, followed by applying matte black spray paint to create a random pattern by the overspray method, as shown in panel (a) of FIG. 47, illustrating a field of view with speckle pattern on specimen with 1=120 mm, w=12.7 mm. The region of interest (colored area) for strain measurement is shown in panel (b) of FIG. 47.

For stereo-DIC measurement, two 4.1 Mpx (2048 px×2048 px) cameras and 35 mm fixed focal length lenses were used. The image acquisition rate was 2 Hertz. The resolution was 60 m to 70 m per pixel and the size of the dark speckles was about 232 m (3.4 px), measured via the line intersection method. The dark/bright ratio of the sample was nearly one (54:46). For data analysis, the chosen subset size was 15 px and the step size was 6 px. The reference image has been taken at a force F=0 kN while the specimen has only been clamped by the top grip. For analysis, engineering strain has been calculated using a commercially available DIC software package.

UD CFRP Hat-Section Dynamic 3-pt Bending Test:

The UD CFRP hat-section studied in the current work was molded with A42 fibers, provided by DowAksa, and thermoset epoxy resin with fiber volume fraction of 50%. The geometry of the dynamic 3 point bending test sample was shown in panel (a) of FIG. 48. The nominal thickness of the hatsection was 2.4 mm with about 0.2 mm thickness of each layer. The hatsection sample was deformed in a hot compress and held for about 5 minutes for curing. The hatsection was made with [0/60/−60/0/60/−60]s layup (noted as 0-60). In order to perform the test successfully, a back plate of the same layup and thickness was glued to the bottom of the hatsection sample with Betamate 4601 glue (Dow), as shown in panel (b) of FIG. 48.

The setup of the dynamic 3 point bending test was shown in panel (c) of FIG. 48. The sample was slightly fixed to the lower roller (diameter 25 mm) with tape in order to allow the rotation at the bottom. An impactor with 25 kg mass and 100 mm outer diameter impacted the hatsection with initial impact velocity of 4.66 m/s. The peak impactor acceleration and the impactor force were recorded for comparison to the numerical predictions.

UD CFRP ICME Modeling Process

In the ICME framework, the CFRP structure is modeled in a bottom-up fashion, as introduced in FIG. 44. At the microscale, UD CFRP microstructure is modeled as RVEs. RVEs are compressed into the microstructure database, which contains ROMs for all RVEs. ROMs can be fed into an arbitrary macroscale FE model composed of 3D stress state elements, such as the brick element or the thick-shell element. The ROMs interact with the macroscale model and enables a multiscale model with information exchange between the microscale and the macroscale. In this section, the multiscale modeling workflow for the ICME process is provided step by step.

ICME Multiscale Modeling Work Flow

For the illustration of ICME multiscale model setup, the UD CFRP coupon specimen geometry described above is used. The UD CFRP coupon FE model mirrors the real UD CFRP coupon. It replicates the off-axial tensile experiment performed as a one-to-one replica. The FE model contains all 12 UD CFRP laminae. Each of these laminae is modeled as a singular layer of thick shell elements with 2 integration points in the thickness direction (Z direction), as shown in FIG. 49. For clarity, the FE model has been magnified by a factor of 2 in the thickness direction. The FE model contains 49,420 elements and 99,480 integration points. The magnified region in FIG. 49 shows four selected thickshell elements and integration points in each element. The UD CFRP microstructure, modeled by the UD CFRP RVE, is assigned to each integration point in order to compute material responses under external loadings.

To illustrate the diversity of the microstructure database, two selected integration points, marked by the red box in FIG. 49, are further magnified for the underneath microstructure. FIG. 50 depicts how two UD RVEs from two neighboring integration points (as indicated by the dotted red box) are modeled using the microstructure database. The database contains four different RVE setups that could potentially be used for the multiscale modeling process. On the right side of FIG. 50, the ellipses around each RVE are used to indicate that there are hidden neighboring RVEs on each side of the RVE due to the assumption of periodic boundary condition (PBC). The first setup assumes perfect bonding between the fiber and matrix, as well as for all laminae. The second setup assumes a weak bond between laminae that can be modeled as a cohesive layer. The third setup assumes an interface region between fiber and matrix, which is modeled as non-zero thickness interphase. The fourth setup assumes there is an interface region between fiber and matrix and that there is weak bond between the laminae. In this work, only the first setup is incorporated due to assumption of perfect bonding.

The RVEs illustrated in FIG. 50 are compressed into UD CFRP ROMs through the three-step offline data compression stage. Details on the offline stage will be provided in later subsections.

UD RVE Modeling

The cured UD CFRP lamina plaque is manufactured by Dow Chemical and the cross-section of the UD CFRP under microscope is shown in FIG. 51. Fibers are shown in lighter color and epoxy is shown in dark color. The fibers are randomly dispersed in the epoxy resin matrix, with overall volume metric fraction of 50%. The fibers are not a perfect circular shape, but rather a bean shape, as shown in FIG. 51. Assuming the bean shape fits into a ellipse domain, then the major axis of the ellipse is measured by the white line segment with length of 7 μm, as shown in the magnified view in FIG. 51. For convenience in modeling of fibers, circular shapes are assumed with a diameter of 7 μm.

RVEs are used to characterize the UD CFRP microstructure. In general, when the RVE is large enough, typically ten times of the fiber diameter, the random distribution of fiber does not affect the RVE responses significantly.

To generate the UD RVE, the fiber cross-section is simplified as a circular shape with a diameter of 7 μm, and the fiber is assumed to be perfectly straight. The cross-section of the UD RVE can be modeled in a 2D fashion, where the Monte Carlo method is used to pack circles randomly in a 2D domain until the target fiber fraction is met. If part of any fiber lies outside four edges of the RVE, that part will reappear in the opposite directions to ensure periodic distribution of all fibers so the RVE complies with the PBC assumption. The generated 2D mesh is then discretized by square pixels. The algorithm flow chart for generating a 2D mesh of the UD RVE is given in the FIG. 52 below. The RVE algorithm discussed above is suitable for UD RVEs with fiber volume fraction around 50%. For higher volume fractions, specialized algorithms are needed. This is beyond the scope of current work and is not discussed in detail. Finally, the 2D mesh is extruded by assigning thickness to all pixels to generate a 3D voxel mesh.

The generated UD RVE is given in FIG. 53. The RVE has a resolution of 600 by 600 by 100 cubic voxels with voxel edge length of 0.14 μm. 93 fibers are generated in the RVE. Fibers are assumed to be perfectly straight and of circular shape.

To utilize UD RVE for compute stress responses on-the-fly in a macroscale FE model, the ROM technique is used to compress RVEs into a microstructure database. The ROM process is given in following sections.

A DNS of RVE transverse tensile loading is also performed. The DNS is used to verify the efficacy of the ROM, which is supposed to produce accurate results compared with DNS solution. Fiber and matrix properties are following data given above

Reduced Order Modeling of UD RVE

The aforementioned UD RVE contains significant DOFs for a single RVE run due to the fine mesh resolution. To model the UD CFRP structure with UD RVEs, the computational cost is not affordable due to the costly RVE computation. Instead, the ROM technique is applied to the UD RVEs to generate their ROMs. The ROM, in theory, will reduce the computational cost significantly compared to RVE computation. The SCA approach is used to generate the ROMs for the UD RVEs. All necessary derivations for SCA are provided in Appendix A. In this subsection, we will focus on illustrating the 3-step offline stage computation and the online prediction stage.

Offline: The offline stage starts with a high fidelity RVE discretized by a voxel mesh. The strain concentration tensor A(x) links macroscopic strain applied on the RVE to each voxel through the following relationship:


εm(x)=A(x):εM  (3-2)

where εm is the microscopic strain at any voxel in the RVE and εM is the applied macroscopic strain of the RVE. A(x) is the well-known strain concentration tensor. Under the Voigt notation, εm and εM are both 1 by 6 vectors. This means A(x) is a 6 by 6 matrix. A(x) can be computed by applying six orthogonal loading conditions where εM has only one non-zero component at a time. This would allow A(x) to be computed one column at a time, and six loading conditions can provide all 36 components of A(x).

Once A(x)s for each voxel are computed, unsupervised learning can be applied to all A(x)s within the RVE to perform clustering. This process will compress the original RVE made of many voxel elements into several clusters. For UD RVE with fiber and matrix phases, fiber and matrix phases are decomposed separately. Number of clusters in fiber and matrix phases are denoted as Kf and Km, respectively. It is convenient to define Kf+Km=K. The clustering process for UD RVE setup one and two, as depicted in FIG. 50, is given in FIG. 54. The data compression process is performed using an unsupervised learning method, such as K-means clustering.

Interaction tensors DIJ must be computed between all cluster pairs. Once interaction tensors are computed, it is possible to solve for cluster-wise strain increments by solving the discretized Lippmann-Schwinger equation.

Online: The online stage involves solving the following residual form in Eq. (3-3). Details of the SCA online stage is given in Appendix A.


rI=−ΔεM+ΔεIJ=1K[DIJ:(ΔσJ−C0:ΔεJ],I=1,2,3, . . . ,K  (3-3)

where rI is the residual of strain increment on each cluster. The residual can be minimized by first linearizing the Eq. (3-3) with respect to ΔεI and solve for ΔεI that minimizes rI.

Note that in the present multiscale modeling scheme, ROMs are deployed on all integration points in the FE mesh of the composite laminate structure. At each integration point, the macroscopic strain increment ΔεM is provided by the FE solver and the cluster-wise stress and strain responses are solved with Eq. (3-3). The homogenized RVE stress increment, denoted as ΔσM, is returned back to the FE solver.

DNS Vs. Reduced Order Model

Two sets of ROMs of the UD RVE are generated with a different number of clusters: 1) 2 clusters in the fiber phase (Kf=2) and 8 clusters in the matrix phase (Km=8); 2) 16 clusters in the fiber phase (Kf=16) and 16 clusters in the matrix phase (Km=16).

To verify that the ROM will produce satisfactory results, a transverse tensile test with maximum 0.02 strain magnitude is performed. The same loading is applied to two ROMs as well. Stress and strain curves for all three cases are plotted in FIG. 55. The DNS result is plotted with a 95% confidence interval. The results of first ROM and second ROM mentioned earlier are denoted as “SCA, 8 cluster” and “SCA, 16 cluster” respectively. From the results shown in FIG. 55, it is clear that both ROMs converge to the DNS solution. To save computational cost in the concurrent multiscale modeling of the UD CFRP, the ROM with 8 clusters is used in the following structural level models for balanced accuracy and efficiency.

As mentioned above, a paraboloid yielding function is implemented to consider different tensile and compressive behavior of the epoxy matrix. The matrix damage is modeled using a paraboloid epoxy damage model.

The UD ROM utilizes fiber and matrix to compute UD CFRP's material responses efficiently. In the UD structure modeling, an equivalent damage model is applied to all ROMs to simulate the damage of RVEs. The damage of the RVE at each integration point will reduce load carrying capacity of each element in the macroscale model. When the damage exceeds 0.5, the integration point will lose load carrying capacity. The macroscale element will fail when all integration points in the element have lost load carrying capacity.

UD CFRP Coupon Off-Axial Tensile Simulation Model Setup

With all aforementioned information available, the UD CFRP Coupon specimen multiscale model is illustrated in FIG. 56. The applied boundary condition is shown in FIG. 56. In the experiments, both tab sections are tightly clamped with a pressure of 4 MPa. The same clamping pressure is applied to both tab sections of the FE model. Moreover, the two surfaces of the upper tab section are fixed in the y direction, but the two surfaces of the bottom tab section are allowed to move in the y direction. A displacement towards the negative y direction is applied to the bottom tab section so the coupon is extended, allowing the tensile test to be repeated. One can see the FE coupon specimen model preserves most of the experimental conditions. This aligns with the purpose of ICME modeling, where a real-world part is modeled and analyzed with as many details as possible.

UD CFRP Dynamic 3-Point Bending Model Setup

The 3-point bending model is the second test case for validating the efficacy of the proposed framework of UD CFRP. A similar approach to the UD CFRP coupon model is adopted. The model of the UD CFRP hat-section is shown in FIG. 57, where a [0/60/−60/0/60/−60]s layup is used. As shown in FIG. 57, the 12 layers of the UD laminate structure are modeled explicitly with thick shell elements. Such a setup allows for the capturing failure of individual lamina under the effect of the impactor.

Up to this point, all necessary modeling steps of the UD CFRP structure are finished. The concurrent multiscale modeling framework is applied to test problems mentioned above. All test problems are modeled with realistic geometries, hence the FE model utilizes the same boundary conditions given in both tests. The experimental test data is used for validating the proposed UD CFRP concurrent multiscale modeling framework. The prediction capability of the framework is examined against experiments conducted delicately. The results of both test problems and associated discussion are given in the next section.

Results and Discussion

Concurrent simulation results of the off-axial coupon tensile test and the 3-point bending test are discussed in this section. The same UD RVE shown above is used in both cases due to same fiber volume fraction.

UD CFRP Coupon Off-Axial Tensile Simulation

In this section, the concurrent multiscale 10° UD off-axial coupon specimen tensile simulation results are presented and compared against experimental results. A loading rate of 0.0167 mm/s gradually extends the coupon sample in the downwards direction. During the deformation process, a 10° stress band is formed as one can see in the FIG. 58. The UD CFRP coupon off-axial tensile simulation is used to validate that the proposed multiscale modeling framework for UD CFRP material predicts the system behavior with good accuracy.

The comparisons between multiscale coupon simulation and the test data are made for: 1) normal stress vs. normal strain; 2) y direction displacement; 3) y direction strain. Based on the comparisons, two purposes are addressed: 1) To validate the multiscale multiscale modeling framework for the UD CFRP material; 2) To demonstrate that the present UD CFRP multiscale model has considerable prediction capability.

For the FE model, the normal stress is computed using the reaction force computed at the gauge cross-section area divided by the area of the original coupon cross-section. The reaction force of the cross-section near the top tab region of the coupon was recorded during the simulation. The reaction force is then used to compute the normal stress of the multiscale model. The change of the gauge length was used to compute normal strain. Normal stress versus strain of the multiscale model is plotted as blue dots in FIG. 58.

Comparing the stress and strain curves from the multiscale model and the experimental data, a good match is observed. The prediction has same trend as the experimental data, as shown in Table 3-3. The predicted maximum stress is 404.81 MPa, which is close to 395.64 MPa reported from the experimental data. In addition, the maximum strain predicted by model is 0.011, which again is in a good match with experimental measured value of 0.012. Both predicted quantities of interest are within 10% deviation from experimental measurements, meaning the UD CFRP concurrent multiscale model is validated against the experimental data.

TABLE 3-3 Predicted maximum normal stress and strain of the off-axial UD coupon sample Maximum Normal Stress Maximum Normal (MPa) Strain Experiment 395.64 0.012 Prediction 404.81 0.011 Difference 2.32% 8.33%

FIG. 59: Contour of a) Y displacement and b) Y strain field. The applied displacement on prediction and DIC is 0.9031 mm. In the displacement plots a), two black arrows measure the vertical distance between fringes from −0.250 mm and −0.700 mm and the difference is 3.95%. In the gray scale strain contour b), the predicted strain field is comparable to the DIC one. The difference the predictions and the DIC images is caused by microstructure variations in the real UD CFRP material, which can cause strain concentration in the real sample.

The y direction displacement and strain fields are validated with experimental results, as shown in panels (a)-)(b) of FIG. 59, respectively. The black arrows in panel (a) of FIG. 59 represent the vertical distance between fringes from −0.250 mm and −0.700 mm. The prediction made by the multiscale model is 80.26 mm, with a 3.95% difference to the DIC measurement of 83.57 mm. As shown in panel (b) of FIG. 59, the y direction strain yield contour predicted is quite similar with the DIC measurement. Both strain field contours show a clear band across the middle of the coupon. The DIC strain field is highly non-uniform due to unavoidable microstructure variation during manufacturing process. Still, the results are encouraging since the predicted displacement and strain fields agree with experimental results.

Moreover, the predicted crack formation and the actual coupon crack formation are depicted in panels (a)-)(b) of FIG. 60, respectively. In both panels (a)-(b) of FIG. 60, the crack propagates all the way across the coupon gauge section. The pattern of predicted crack is not an exact replica of the experimental results because the numerical model has yet to take local microstructure variation into account, as depicted in FIG. 51. In the future work, the local microstructure variation will be considered in the modeling process to address the uncertainty effect on the macroscopic performance. Nonetheless, the ICME model is able to provide an accurate prediction of the maximum stress and strain of the coupon sample, as reported in Table 3-3, as well as a prediction of the failure pattern. The prediction made by the ICME model is consistent with the experimental observation, which provides confidence for the predictivity of the proposed multiscale modeling scheme for the ICME process.

The ICME multiscale modeling scheme has been validated by the experimental data discuss above. However, the capability of the proposed multiscale modeling scheme is beyond providing accurate prediction of the macroscale quantities. It also provides detailed microstructure evolution of UD CFRP structure for studying the root cause of the failure in the UD coupon.

FIG. 61 illustrate the local UD CFRP damage process using three snapshots. The magnified views of the region marked in red are shown in panels (a)-(c) of FIG. 61 at y displacement of 1.40 mm, 1.41 mm, and 1.42 mm. The UD RVEs representing elements marked in black are shown in panels (d)-(f) of FIG. 61. In panel (e) of FIG. 61, all three RVEs are experiencing matrix damage, suggesting the potential cause of coupon failure. In panel (b) of FIG. 61, the bottom element has a stress contour turning to blue-green from green, suggesting reduced load carrying capacity. This is due to aggravated matrix failure of the middle and bottom RVEs as shown in panel (e) of FIG. 61. In panel (c) of FIG. 61, the middle and bottom elements have been marked as failed due to the loss of load carrying capacity, whereas the top one is still capable of carrying load. However, as shown in panel (f) of FIG. 61, the top RVE is also experiencing severe matrix damage, which means this RVE will fail shortly. The UD RVE microstructure evolution in FIG. 61 provides valuable information in the understanding of microscale damage process. Moreover, the UD RVE stress and damage evolution are captured simultaneously as the macroscopic UD coupon simulation. In the future, in-situ monitoring techniques can be combined with the existing modeling capability to further validate microstructure failure process of the UD

Using detailed microstructure evolution information illustrated in FIG. 61, a path towards UD CFRP design is discovered. For example, one can design matrix strength to avoid matrix damage at small loading magnitude. Or, one can design interphase properties and incorporate the interphase region into the model to examine the interphase effect and the delamination behavior of the UD CFRP.

UD 3-Point Bending Model

To further illustrate the efficacy of the concurrent scheme, the UD 3-point bending model concurrent simulation is also performed. The fractured hat-section of the simulation and experiment is shown in FIG. 64. The UD 3-point bending simulation is used to examine the applicability of the proposed scheme to an arbitrary UD CFRP structure. Two quantities are chosen to make the comparison: peak load on the impactor and the peak impactor acceleration. Due to severe vibration of the hat-section in the 3-point bending simulation, the matrix has been simplified as an elastic material with brittle failure, where the failure strength is set to compression strength as defined in Table 3-2.

After the impact, the hat-section from the numerical prediction and experimental result are plotted in panels (a)-(b) of FIG. 62, respectively. The yellow ellipse marked damaged zones on the hat-section. It can be observed that the impact push material inwards, cause delamination of the hat-section on both side-walls. The similar trend observed in the numerical model and the experimental result further suggests the multiscale model can provide a good prediction of the failure pattern on the UD structure. The ICME process can be applied to predict the responses of a UD structure under complex loading conditions.

In addition to the good match of the UD structure deformation and failure, quantitative comparisons for peak load on the impactor and peak acceleration of the impactor are reported in Table 3-4 with comparison with the experimental data. A reasonable match between prediction and experimental data can be seen. Specifically, the relative differences of peak load and peak impactor acceleration are 8.21% and 2.82%, respectively. Both predictions are within 10% deviation from the experimental data, providing confidence that the macroscopic performance indices can be predicted with the ICME framework.

TABLE 3-4 Impactor peak load and acceleration Peak Impactor Peak Load(N) Acceleration (m/s2) Experiment 10328 0.39 Prediction 9480 0.379 Difference 8.21% 2.82%

In panel (a) of FIG. 62, the failure of the hat-section is depicted at the macroscale. Underneath the complex macroscopic structure evolution, local RVE responses are captured in a concurrent fashion and provide extra information for the microscale evolution in the hat-section. To illustrate the von Mises stress evolution in the UD microstructure, RVEs representing three columns of elements shown in FIG. 63 are visualized. The three columns of elements are shown in the magnified view on the right of FIG. 62. Each column contains 12 layers of UD laminae, where each lamina is visualized with 1 UD RVE. The fiber orientation on each UD laminate is color coded as shown in FIG. 63. All UD RVEs are aligned with the fiber orientation as suggested by the color code, following the lamina orientation. All UD layers are counting from layer 1 to layer 12 in the top-down fashion.

The hat-section deformation and von Mises stress contour under the impactor are shown in the upper half of FIG. 64. Three different snapshots were taken with the impactor and supports hidden at: upon impact; Impactor displacement of 4.85 mm; Impactor displacement of 6.85 mm. In the upper half of panel (a) of FIG. 64, the impactor contacts the hat-section and caused immediate stress concentration at the contact region. In panels (b)-(c) of FIG. 64, it can be observed that the top and two sidewalls of the hat-section bend inwards after the hat-section contacts the impactor. The impactor pushes the middle of the UD CFRP hat-section inwards, simulating the scenario where the composite structure is under the external loading condition.

Corresponding RVE von Mises stress contours are shown in the lower half of panels (a)-(c) of FIG. 64. It can be seen that UD RVEs can be used to investigate the inter-laminate stress distribution of the UD microstructure. Due to different fiber orientations in the hat-section laminate structure, the von Mises stress magnitude varies from layer to layer across the thickness direction, which might lead to earlier failure for those layers with larger stress magnitude. The RVE plots shown in panel (b) of FIG. 64 suggested that the second and third in location 1 has much higher stress magnitude comparing with those in other layers. In panel (c) of FIG. 64, those two layers have reached failure as the stress magnitude are zero. This suggested that the multiscale modeling has great potential in predicting microstructural evolution in a complex UD CFRP laminate structure. Such information can be used to examine various UD CFRP laminate designs using the virtual verification capability to examine macro and micro material performance. This can assist the CFRP design process by eliminating designs that under-perform certain performance indices, or designs with undesired microstructural evolution pattern.

For UD CFRP materials, understanding the microstructure failure mechanism provides valuable information in improving CFRP design. Similar to the UD CFRP coupon model, the microstructure evolution of the hat-section is captured and illustrated in the three-snapshot view in FIG. 65 with different impactor displacement. In FIG. 65, two UD CFRP RVEs representing two marked regions are visualized. It can be seen that the RVE in the upper layer begins to damage while the RVE in the lower layer stays intact when dy=5.91 mm, as shown in panel (a) of FIG. 65. When the upper layer fails, the neighboring region will collapse towards the newly formed empty region and the neighbor elements will compress the lower layer. Soon, the lower layer fails as well when dy=6.13 mm, as shown in panel (b) of FIG. 65. Eventually, both layers fails as shown in panel (c) of FIG. 65. This three snapshots illustrate the ability of investing detailed microstructure evolution of UD CFRP for a structural level simulation. Such information can provide guidance for design of CFRP structure against local damage to improve the structural performance.

In this subsection, a one-to-one replica of the UD CFRP Coupon 10° off-axial tensile test multiscale simulation and the UD CFRP hat-section 3-pt bending test are resolved using the proposed multiscale modeling framework. Both models are validated against experimental data for validation of the framework. Predictions made by the numerical counterparts are all within 10% difference compared with experimental data. The agreement shows the current work can be used for prediction of other UD CFRP laminate structure. Moreover, the concurrent capture of microstructure evolution provides microstructure evolution history for any location in the FE model of the UD laminate. This allows researchers to look at the detailed microstructure evolution, which is hard to capture experimentally, for the cause of failure at the structural level. New microstructure can be then designed to sustain the loading and improve the overall structural performance.

Briefly, the exemplary study introduces a predictive and efficient ICME multiscale modeling framework for UD CFRP materials. The main workflow of the framework is explained in detail, and experimental validations are provided. The predictive framework links UD CFRP microstructures to structural level models for accurate prediction of the structural performance. Two sample cases studies, the UD off-axial tensile test and the UD hat-section dynamic three-point bending test, are presented using the proposed ICME modeling framework. The predicted performance indices are validated against experimental data confirming a good agreement. Microstructure evolution in the UD structure is captured by UD CFRP RVEs, which reveal microstructural evolution, including stress contour and matrix damage. The ICME framework is general and can be applied to other Fiber Reinforced Polymer (FRP) systems, such as glass fiber reinforced polymer, for structure performance prediction. Along with the microstructure information, the work presented in this example should provide guidance to existing experimental based composites design workflows to accelerate the design process.

Future work of the present multiscale modeling framework for UD CFRP includes: 1) Incorporation of the interphase in the UD RVE for fiber-matrix debonding; 2) Consideration of microstructure uncertainties, such as fiber misalignment and fiber volume fraction, for quantitative measurement of the microstructure effect. 3) Extension to other composite material systems.

Example 4 Data-Driven Self-Consistent Clustering Analysis of Heterogeneous Materials with Crystal Plasticity

To analyze complex, heterogeneous materials, a fast and accurate method is needed. This means going beyond the classical finite element method, in a search for the ability to compute, with modest computational resources, solutions previously infeasible even with large cluster computers. In particular, this advance is motivated by composites design. Here, we apply similar principle to another complex, heterogeneous system: additively manufactured metals.

The complexities and potential benefits of metal additive manufacturing (AM) provide a rich basis for development of mechanistic material models. These models are typically based on finite element modeling of metal plasticity. Powder-bed AM uses a high-power laser or electron beam to melt powder layer-by-layer to produce freeform geometries specified by 3D model files. This approach removes the need for special tooling, allowing for rapid and customized part and product realization. It introduces new possibilities for topological and material optimization, but these tasks require a high degree of knowledge and ability to apply that knowledge, viz. control the build conditions sufficiently. The process involves intense and repeated localized energy input, which results in inhomogeneous, anisotropic, location-dependent material properties with complex microstructures.

The perpetual challenge in multiscale modeling is predicting macroscopic behavior from microstructure conformation and properties in both an efficient and accurate manner. Analytical approaches such as the rule of mixtures and other micromechanics methods are very efficient but lose accuracy particularly when dealing with irregular morphologies and nonlinear properties. In contrast, direct numerical simulations (DNS), offer high accuracy at the expense of prohibitive computational costs to the point where they are inapplicable to concurrent simulations for material design. Recently, data mining has been introduced into the mechanics community to address the limitations of DNS and analytical methods.

In general, data mining is a computational process of discovering patterns in large data sets. Once extracted, these patterns can be used to predict future events. Machine learning methods are the technical basis for data mining, such as clustering and regression methods. Recently, data mining has also been applied to the modeling of heterogeneous materials. As a start, a raw dataset for learning is usually generated from a priori numerical simulations or informed by experiments. Depending on the type of the raw dataset, current data-driven modeling methods can be mainly divided into macroscopic and microscopic approaches.

In macroscopic approaches, the input data are usually material properties of each constituent, loading conditions and statistical descriptors that represent the geometry of the microstructures, while the output data are macroscopic mechanical properties from direct numerical simulations (DNS). However, the accuracy and smoothness of the prediction of macroscopic approaches is limited by a lack of microscopic information. For example, the localized plastic strain fields, critical for plasticity and damage prediction theories, cannot be well represented by their field averages.

To address this problem, microscopic approaches collect data at each discretization point in the DNS. Two methods for making predictions based on gathered local data are worth highlighting: (1) non-uniform transformation field analysis (NTFA) and (2) variants of the proper orthogonal decomposition (POD) method. For both approaches, the predictions under a loading condition are obtained by linear combinations of a finite number of RVE modes from previously completed simulations under various load conditions. Linear combination of eigenmodes is well established, but extra effort is required for the interpolation for nonlinear materials. For NTFA, specific evolution laws of internal variables have to be assumed for each mode of the inelastic field. For POD-based methods, extensive simulations a priori are needed in order to guarantee the robustness of the interpolation under arbitrary loading conditions. However, this still results in an overall decrease in computational cost.

One of the objectives of this exemplary study is to present a grain-level crystal plasticity model to capture local microstructures such as those that occur in AM, e.g., voids and columnar grains. In one embodiment, a recently introduced material modeling approach, based on the theories of data mining and originally developed for composites, is used to vastly increase the speed of these simulations. This allows for higher detail or larger regions of interest, both of which are desirable for predicting damage and fatigue initiation within a part.

The exemplary study is based on a two-stage approach that uses clustering and subsequent analysis of deformation and can account for heterogeneous material behavior with high accuracy and speed, which is called SCA. It is a data-driven method designed to reduce the computational degrees of freedom (DOFs) required for predicting macroscopic behaviors of heterogeneous materials, while local information is retained based in part on clustering near features that induce large stress gradients. The basic idea is to solve the equilibrium equation, not at every material point, but on clusters of material points with similar mechanical responses by assuming that local variables of interest (e.g., elastic strain, plastic strain and stress) are uniform in each of these clusters. The two stages of the method are: offline (training or cluster) and online (prediction).

During the offline stage, material points were grouped into clusters using data mining techniques (such as k-means clustering) based on mechanical similarity. To conduct the online computation, the equilibrium equation was written in an integral form using the Green's function, known as the Lippmann-Schwinger equation. This equation was solved for each cluster using a self-consistent scheme to ensure the accuracy, where the reference stiffness was updated iteratively to be consistent with the macroscopic effective stiffness. The major advantage of SCA is that the DOFs are greatly reduced compared to DNS while retaining both local and global response information.

Here we extend this method to be applicable to crystal plasticity (CP), a class of computational plasticity problems specifically formulated to capture the deformation mechanics of crystalline solids, based on the material microstructure. Anisotropic material models such as CP have been derived for both macroscale problems, such as predicting earing during deep drawing, and microscale problems, such as the deformation of nanowires.

SCA Framework

Offline Stage: Mechanistic Material Characterization

Grouping material points with similar mechanical behavior into a single cluster is performed by domain decomposition of material points using clustering methods. First, the similarity between two material points is measured by the strain concentration tensor A(x), which is defined as


εmicro(x)=A(x):εmacroinΩ,  (4-1)

where εmacro is the elastic macroscopic strain corresponding to the boundary conditions of the Representative Volume Element (RVE), and εmicro(x) is the elastic local strain at point x in the microscale RVE domain Ω. For a 2-dimensional (2D) model, A(x) has nine independent components, requiring a set of elastic direct numerical simulations (DNS) under three orthogonal loading conditions to uniquely define. For a 3-dimensional (3D) model, A(x) has 36 independent components which are determined by DNS under six orthogonal loading conditions. Once the strain concentration tensor is computed, it is independent of the loading conditions for a linear elastic material, and its Frobenius norm is an invariant under coordinate transformation.

For overall responses of nonlinear plastic materials, we have demonstrated that the elastic strain concentration tensor is a good offline database. However, if the local response is of more interest, the elastic strain concentration tensor often does not provide high enough cluster density near the high strain concentration region. In polycrystalline material with crystal plasticity, all the crystals deform uniformly in the elastic regime, providing no effective data for computing the strain concentration tensor. In these cases, one can choose other types of material responses to construct the offline data, and thus achieve adequate resolution at the region of interest. For example, we choose the plastic strain tensor from DNS calculations for clustering when local plasticity information is required, such as for predicting fatigue initiation. We show how the choice of different material responses affect overall response prediction.

The k-means clustering method is used to group data points based on a grouping metric. For present, let us consider this to be the strain concentration tensor A(x). Since all the material points in a cluster are assumed to have the same mechanical response, the number of the degrees of freedom is reduced from, e.g., the number of elements in a FEM simulation to the number of clusters. Note that clusters are formed based on the strain concentration tensor and thus do not need to be spatially adjacent to each other.

A primary assumption associated with the domain decomposition is that any local variable β(x) is uniform within each cluster. Globally, this is equivalent to having a piece-wise uniform profile of the variable in the RVE:


β(x)=ΣI=1kβIχI(x),  (4-2)

where βI is the homogeneous variable in the Ith cluster, and k is the total number of clusters in the RVE. The domain of the Ith cluster ΩI is distinguished by its characteristic function χI(x), which is defined as

χ I ( x ) = { 1 if x Ω I 0 otherwise ( 4 - 3 )

This piecewise uniform approximation in Eq. (4-2) enables us to reduce the number of degrees of freedom for the Lippmann-Schwinger equation, which is solved in the following online stage. After the domain decomposition based on a prior DNS, the remaining task in the offline stage is to pre-compute the interaction tensors between all the clusters.

In the discretized/reduced Lippmann-Schwinger equation, we can utilize the piecewise uniform assumption to extract the interaction tensor DIJ, which represents the influence of the stress in the Jth cluster on the strain in the Ith cluster. In an RVE domain Ω with periodic boundary conditions, the interaction tensor can be written as a convolution of the Green's function and the characteristic functions defined in Eq. (4-3):

D IJ = 1 c I Ω Ω Ω χ I ( x ) χ J ( x ) Φ 0 ( x , x ) dx dx , ( 4 - 4 )

where cI is the volume fraction of the Ith cluster and |Ω| is the volume of the RVE domain. Φ0(x,x′) is the fourth-order periodic Green's function associated with an isotropic linear elastic reference material and its stiffness tensor is C0. Specifically, this reference material is introduced in the online stage as a homogeneous media to formulate the Lippmann-Schwinger integral equation. With the periodicity of the RVE, Φ0(x,x′) takes the following form in the Fourier space,

Φ ^ 0 ( ξ ) = 1 4 μ 0 Φ ^ 1 ( ξ ) + λ 0 + μ 0 μ 0 ( λ 0 + 2 μ 0 ) Φ ^ 2 ( ξ ) , with ( 4 - 5 ) Φ ^ i j k I 1 ( ξ ) = 1 ξ 2 ( δ i k ξ j ξ l + δ il ξ j ξ k + δ j t ξ i ξ k + δ j k ξ i ξ l ) ( 4 - 6 ) Φ ^ ijkl 2 ( ξ ) = - ξ i ξ j ξ k ξ l ξ 4 , ( 4 - 7 )

where ξ is the coordinate in Fourier space corresponding to x in real space, and δij is the Kronecker delta function. λ0 and μ0 are Lamé constants of the reference material. The expression of {circumflex over (Φ)}ijkl0(ξ) is not well defined at frequency point ξ=0. However, by imposing the boundary conditions for deriving the Green's function, a uniformly distributed polarization stress field will not induce any strain field inside the RVE. As a result, we have


{circumflex over (Φ)}ijkl0(ξ=0)=0.  (4-8)

Based on Eq. (4-5), the convolution term in the spatial domain in Eq. (4-4) can be translated into a direct multiplication at each point in the frequency domain using a Fourier transformation,


ΦJ0(x)=ωΩχJ(x′)Φ0(x,x′)dx′=−1({circumflex over (χ)}(ξ){circumflex over (Φ)}0(ξ)).  (4-9)

As we can see from Eq. (4-6), {circumflex over (Φ)}1(ξ) and {circumflex over (Φ)}2(ξ) are independent of the material properties, so that they can be computed once, in the offline stage. If the reference material is changed in the self-consistent scheme during the online stage only the coefficients relating to the reference Lamé constants in Eq. (4-5) need to be updated. For RVEs with microstucture size close to the RVE size or even with a connected microstructure network, such as a woven composite, a correction of DIJ is needed to satisfy the boundary conditions on the RVE.

Currently, we have applied the SCA offline stage to 2D and 3D materials with uniform (regular hexahedral or “voxel”) meshes, so that the Fast Fourier transformation (FFT) method can be used for efficiently computing Eq. (4-9). Although the domain decomposition is based on a specific selection of properties for each material phase in the offline stage, the same database can be used for predicting responses for new combinations of material constituents in the online stage.

Online Stage: Self-Consistent Lippmann-Schwinger Equation

The equilibrium condition in the RVE can be rewritten as a continuous Lippmann-Schwinger integral equation by introducing a homogeneous reference material,


Δε(x)+∫ΩΦ0(x,x′):[Δσ(x′)−C0:Δε(x′)]dx′−Δε0=0,  (4-10)

where Δε0 is the far-field strain increment controlling the evolution of the local strain. It is uniform in the RVE. The reference material is isotropic and linear elastic. Its stiffness tensor C0 can be determined by the two independent Lamé parameters λ0 and μ0,


C0=f00)=λ0I⊗I+μ0II.  (5-11)

where I is the second-rank identity tensor, and II is the symmetric part of the fourth-rank identity tensor. The strain and stress increments are Δε(x) and Δσ(x). By averaging the incremental integral equation, Eq. (4-10), in the RVE domain Ω, we have

1 Ω Ω Δ ɛ ( x ) dx + 1 Ω [ Ω Φ 0 ( x , x ) dx ] : [ Δ σ ( x ) - C 0 : Δɛ ( x ) ] dx - Δ ɛ 0 = 0. ( 4 - 12 )

Imposing by the boundary conditions for deriving the Green's function, Eq. (4-8) can be equivalently written as


ΩΦ0(x,x′)dx=0.  (4-13)

Substituting Eqs. (4-13) into (4-12) gives

Δ ɛ 0 = 1 Ω Ω Δ ɛ ( x ) d x , ( 4 - 14 )

which indicates that the far-field strain increment is always equal to the ensemble averaged strain increment in the RVE. In order to solve Δε(x) in the integral equation, Eq. (4-10), constraints are needed from the macroscopic boundary conditions. The macro-strain constraint can be written as

1 Ω Ω Δ ɛ ( x ) d x = Δ ɛ _ m a c r o or Δɛ 0 = Δ ɛ _ m a c r o , ( 4 - 15 )

where Δεmacro is the macroscopic strain increment applied on the RVE. Similarly, the macro-stress constraint can be related to the macroscopic stress σmacro,

1 Ω Ω σ ( x ) dx = σ _ macro ( 4 - 16 )

For more general cases, mixed constraints (e.g., strain constraint in the 11-direction and stress constraints for the rest) can also be formulated.

As the full-field calculations (e.g., FFT-based method) of the continuous Lippmann-Schwinger equation may require excessive computational resources, we will perform the discretization of the integral equation based on the domain decomposition in the offline stage. With the piecewise uniform assumption in Eq. (4-2), the number of degrees of freedom and the number of internal variables in the new system can be reduced. After decomposition, the discretized integral equation of the Ith cluster is:


ΔεIJ=1kDIJ:[ΔσJ−C0:ΔεJ]−Δε0=0,  (4-17)

where ΔεJ and ΔσJ are the strain and stress increment in the Jth cluster. The interaction tensor DIJ is defined in Eq. (4-4), which is related to the Green's function of the reference material. After discretization, the far field strain is still equal to the average strain in the RVE,


Δε0I=1kcIΔεI.  (4-18)

The macroscopic boundary conditions also require discretization. For instance, the discrete form of the macro-strain constraint can be written as


ΣI=1kcIΔεIεmacro or Δε0εmacro  (4-19)

In the new reduced system, the unknown variables are the strain increments in each cluster ΔεI. Significantly fewer clusters than FE nodes means that the ROM is much faster to solve. In a general case such as plasticity, the stress increment ΔσJ is a nonlinear function of its strain increment ΔεJ, and Newton's method is used to solve the nonlinear system iteratively for each increment.

The solution of the continuous Lippmann-Schwinger equation (4-10) is independent of the choice of the reference material C0. This is because the physical problem is fully described by the equilibrium condition and the prescribed macroscopic boundary conditions. However, the equilibrium condition is not strictly satisfied at every point in the RVE for the discretized equations because of the piecewise uniform assumption, and the solution of the reduced system depends on the choices of C0. This discrepancy can be reduced by increasing the number of clusters used, at the cost of increased computational cost due to the increased degrees of freedom.

To achieve both efficiency and accuracy, we propose a self-consistent scheme in the online stage, which retains accuracy with fewer clusters. In the self-consistent scheme, the stiffness tensor of the reference material, C0 is approximately the same as the homogenized stiffness tensor C,


C0→C.  (4-20)

Material non-linearity generally makes it impossible to determine an isotropic C0 exactly matching C. Here we propose two types of self-consistent schemes to approximate Eq. (4-20): 1) linear regression of average strain increment Δε and stress increment Δσ (or regression-based scheme) and 2) isotropic projection of the effective stiffness tensor C (or projection-based scheme).

Regression-Based Self-Consistent Scheme

In the regression-based scheme, the self-consistent scheme is formulated as an optimization problem: the goal is to find an isotropic C0 that minimizes the error between the predicted average stress increments. The inputs of the regression algorithm are the average strain increment a and stress increment Δσ, which are computed as


ΔεI=1kcIΔεI and ΔσI=1kcIΔσI  (4-21)

The objective of the regression-based scheme is to find the λ0 and μ0 of the reference material by computing

{ λ 0 , μ 0 } = arg min { λ , μ } Δ σ ¯ - f ( λ , μ ) : Δ ɛ ¯ 2 . ( 4 - 22 )

where ∥Z∥2=Z:Z for an arbitrary second-order tensor Z. The function ƒ(λ′,μ′) can be expressed as


ƒ(λ′,μ′)=λ′I⊗I+μ′II.  (4-23)

where I is the second-rank identity tensor, and II is the symmetric part of the fourth-rank identity tensor. Equivalently, the cost function g(λ′,μ′) of the optimization problem can be written as


g(λ′,μ′)=∥Δσ−ƒ(λ′,μ′):Δε2.  (4-24)

The optimum point is found via the respective partial derivatives of the cost function,

g λ′ λ 0 , μ 0 = 0 and g μ λ 0 , μ 0 = 0 , ( 4 - 25 )

which forms a system of two linear equations in terms of the Lamé constants. The system always has a unique solution except under a pure-shear loading condition, where λ0 is under-determined. In this case, the value of λ0 is not updated. Additionally, g(λ00) vanishes when the effective macroscopic homogeneous material is also isotropic linear elastic.

Although this scheme does not require computing C explicitly, it has two main drawbacks. First, the optimization problem is under-determined for hydrostatic and pure shear loading conditions, forcing one of the two independent elastic constants to be assumed. Second and more important, the modulus of the optimum reference material may be negative for complex loading histories within a concurrent simulation, which is deleterious to the convergence of the fixed-point method.

Projection-Based Self-Consistent Scheme

To avoid the difficulties of the regression-based scheme, we present another self-consistent scheme based on isotropic projection of the effective stiffness tensor C. Through the homogenization, the effective stiffness tensor C of the RVE can be expressed as


C=ΣI=1kcIcICalgI:AI,  (4-26)

where CalgI is the algorithm stiffness tensor of the material in the Ith cluster and is an output of the local constitutive law for the current strain increment in the cluster,

C alg I = Δσ I Δɛ I ( 4 - 27 )

The strain concentration tensor of the Ith cluster AI relates the local strain increment in the Ith cluster ΔεI to the far-field strain increment Δε0,


ΔεI=AI:Δε0,  (4-28)

which can be determined by first linearizing the discretized integral equation (4-17) using CalgI and then inverting the Jacobian matrix. Since C is only required for the self-consistent scheme, the calculation of C can be performed once, after the Newton iterations have converged, to save computational cost.

For a 3D problem, the stiffness tensor of the isotropic reference material C0 can be decomposed as


C0=(3λ0+0)J+0K,  (4-29)

where the forth-rank tensors J and K are defined as


J=⅓(I⊗I) and K=II−J.  (4-30)

Since the two tensors are still orthogonal, we have


J::K=0,J::J=1,K::K=5.  (4-31)

Based on Eq. (4-31), the projection from the homogenized stiffness tensor C to C0 can be expressed as


C0=Ciso=(J::C)J+1/5(K::C)K.  (4-32)

Meanwhile, the Lamé parameters λ0 and μ0 of the reference material can also be determined from the isotropic projection. Since C0 is approximated based on C directly, it is guaranteed to be positive definite the condition of C. Overall, the projection-based scheme can be considered a relaxation of the regression-based scheme.

Summary of the Online Stage

In both schemes, the optimum reference material must be determined iteratively since the values of C in Eq. (4-26) are obtained by solving the reduced system with a previous C0. A fix-point method is employed here. For this method, the convergence of the reference material parameters can be reached in only a few iterations in our experience (i.e., less than five reaches a tolerance of 0.001).

The general algorithm for solving the self-consistent Lippmann-Schwinger equation is as follows:

    • 1. Initial conditions and initialization: set (λ00); {ε}0=0;n=0; {Δε}new=0;
    • 2. For loading increment n+1, update the coefficients in the interaction tensor DIJ and the stiffness tensor of the reference material C0;
    • 3. Start Newton iterations:
      • (a) compute the stress difference {Δσ}new based on the local constitutive law (b) use that compute the residual of the discretized integral equation (4-17): {r}=ƒ({Δε}new,{Δσ}new);
      • (c) compute the system Jacobian {M};
      • (d) solve the linear equation {dε}=−{M}−1{r};
      • (e) {Δε}new←{Δε}new+{dε}; and
      • (f) check error criterion; if not met, go to 3(a);
    • 4. Calculate (λ00) using the regression-based scheme (4-22) or the projection-based scheme (4-32);
    • 5. Check error criterion; if not met, go to step 2;
    • 6. Update the incremental strain and stress: {Δε}n+1={Δε}new, {Δσ}n+1={Δσ}new;
      • Update the index of loading increment n←n+1; and
    • 7. If simulation not complete, go to step 2.

The relative iterative error criterion to the L2 norm of the residual is used. If all the phases of the material are linear elastic, the Newton iteration will converge in one step. Note that if the self-consistent scheme is not utilized for in the calculation, a constant stiffness tensor C0 will be used, which can be chosen to be same as the matrix material. In this case, C0 is not updated, which implies that the interaction tensors DIJ do not need to be updated, and steps 4-5 in the algorithm can also be skipped. Although the algorithm with a constant C0 can save time in terms of finding the optimum C0, the accuracy in predict nonlinear material behavior cannot be guaranteed with a small number of clusters.

Crystal Plasticity Model

In this work, we present an elasto-plastic, anisotropic, heterogeneous plasticity model of the mechanical response of crystalline materials, to be solved in the SCA framework described in above. The mechanical model is an implementation of so-called crystal plasticity (CP) constitutive laws. The finite element scheme used therein to solve the variational form of the equilibrium equations is replaced with the SCA scheme and its FFT-basis. Thus, in some regards this begins to resemble recent CP-FFT formulations, with the addition of an the offline/online decomposition outlined above.

Brief Overview

Crystal plasticity in conjunction with the finite element method (termed CPFEM) has been applied to solve both microscopic and macroscopic problems, following from the early combinations of classical plasticity and the finite element method. It has two primary variants: polycrystal and single crystal plasticity. In the polycrystal formulation, each material point is assumed to represent a collection of crystals such that the overall response of the point is homogeneous. In single crystal plasticity, each material point is assumed to represent a single crystal, or a point in a single crystal, the deformation of which is governed by the particularities of single crystal deformation mechanics (e.g., active slip systems and/or dislocation motion). The former approach is more commonly used for macroscopic problems, where a relatively large solution volume is desired. The later is the focus of this study. There are many versions of crystal plasticity laws in both forms. Here the basic kinematics and constitutive law are discussed below.

Kinematics

The deformation gradient F can be multiplicatively decomposed as:


F=FeFp  (4-33)

where the plastic part Fp maps points in the reference configuration onto an intermediate configuration which is then mapped to a current configuration through the elastic part Fe. Note that physically Fp is associated with the dislocation motion and Fe is a combination of the elastic stretch and rigid body rotation.

The effect of dislocation motion is modeled by relating the plastic velocity gradient {tilde over (L)}p in the intermediate configuration (usually denoted by {tilde over (□)}) to simple shear deformation γ(α):


{tilde over (L)}pα=1Nslip{dot over (γ)}(α)({tilde over (s)}(α)⊗ñ(α))  (4-34)

where ⊗ is the dyadic product, Nslip is the number of slip systems, {dot over (γ)}(α) is a shear rate, {tilde over (s)}(α) is the slip direction, and ñ(α) is the slip plane normal, all for a crystal slip systems (α) in the intermediate configuration. The relationship between {tilde over (L)}p and Fp is given by


{tilde over (L)}p={dot over (F)}p·(Fp)−1.  (4-35)

Constitutive Law

The final task in constructing the crystal plasticity framework is defining the constitutive laws of elasto-plasticity. We choose a basis of the Green-Lagrange strain Ee and Second Piola-Kirchhoff stress Se, from the many conjugate pairs available, which are related by:


Se={tilde over (C)}·Ee=1/2{tilde over (C)}·[(Fe)TFe−I],  (4-36)

where the elastic stiffness tensor {tilde over (C)} is defined in the intermediate configuration.

A phenomenological power law for the plastic shear rate in each slip system given by

γ . ( α ) = γ . 0 τ ( α ) - a ( α ) τ 0 ( α ) ( m - 1 ) ( τ ( α ) - a ( α ) τ 0 ( α ) ) ( 4 - 37 )

is used, where τ(α) is the resolved shear stress, a(α) is a backstress that describes kinematic hardening, {dot over (γ)}0 is a reference shear rate, τ0(α) is a reference shear stress that accounts for isotropic hardening, and m is the material strain rate sensitivity. Shear stress is resolved onto the slip directions with:


τ(α)=σ:(s(α)⊗n(α)),  (4-38)

where σ, s(α) and n(α) are the Cauchy stress, slip direction and slip plane normal respectively, all of which are in the current configuration. The Cauchy stress is given by:

σ = 1 J e [ F e · S e · ( F e ) T ] , ( 4 - 39 )

where Je is the determinate of Fe. The relationship between s(α) and {tilde over (s)}(α) is given by


s(α)=Fe·{tilde over (s)}(α),  (4-40)

and the relationship between n(α) and ñ(α) is given by


n(α)(α)·(Fe)−1  (4-41)

which ensures that the slip plane normal vector remains orthogonal to the slip direction in the current configuration.

The reference shear stress τ0(α) evolves based on the expression:


{dot over (τ)}0(α)=HΣβ=1Nslipqαβ{dot over (γ)}(β)−Rτ0(α)Σβ=1Nslip|{dot over (γ)}(β)|,  (4- 42)

where H is a direct hardening coefficient and R is a dynamic recovery coefficient and qαβ is the latent hardening ratio given by:


qαβ=χ+(1−χ)δαβ  (4-43)

where χ is a latent hardening parameter. The backstress a(α) evolves based on the expression:


{dot over (a)}(α)=h{dot over (γ)}(α)−ra|{dot over (γ)}(α)|,  (4-44)

where h and r are direct and dynamic hardening factors respectively.

A computational crystal plasticity algorithm needs to solve a set of non-linear equations from Eq. (4-33) to Eq. (4-44). Different numerical methods can be used to solve these equations. McGinty and McDowell gave an implicit time integration algorithm for the material law with the finite element method. However, the SCA method uses Fast Fourier Transformation method, CP alorithms have been shown to be effective in this framework as well. Here we simply implement the same crystal plasticity model in our SCA and FEM calculations, albeit with a slight variation in how the deformation gradient, F, is computed.

EXAMPLES

In this section, three example cases probing the capabilities of SCA are implemented with the CP routine described. First, the SCA method is validated for a multi-inclusion system with J2 plasticity. Second, a simple case of a spherical inclusion in a single-crystal matrix is shown. Finally, the complexity of the system is increased by simulating a polycrystalline cube with equiaxed, randomly oriented grains.

Multi-Inclusion System with J2 Plasticity

The SCA method is firstly validated for a multi-inclusion system using a much simpler material law: J2 plasticity. The inclusion phase is elastic with Young's modulus Ei=500 MPa and Poisson's ratio vi=0.19. The matrix phase is elasto-plastic with Em=100 MPa and vm=0.3 in the elastic regime, and it has a von Mises yield surface (J2 plasticity) and a piece-wise hardening law depending on the effective plastic strain εp:

σ Y ( ɛ p ) = ( 0.5 + 5 ɛ p ɛ p [ 0 , 0.04 ) 0.7 + 2 ɛ p ɛ p [ 0.04 , ) MPa . ( 4 - 45 )

The inclusions are identical to each other and the volume fraction of the inclusion phase is equal to 20%. The mesh size for the high-fidelity finite element model is 80×80×80. The clustering results based on the strain concentration tensor A(x) are shown in FIG. 67. The numbers of clusters in the matrix and inclusion are denoted by km and ki, respectively. Note that A(x) has 36 independent components which need to be determined by elastic simulations under 6 orthogonal loadings. Since the volume fraction of the inclusion phase is 20%, we choose ki=┌km/4┐, where ┌□┐ denotes the nearest integer greater than or equal to □.

The stress-strain curves predicted by the regression-based and projection-based self-consistent schemes are given in FIG. 68 for uniaxial tension and pure shear loading conditions, where the solid lines represent the DNS results for comparison. The DNS results are plotted as solid lines for comparison. For both schemes, the predictions converge to the DNS results by increasing the number of clusters in the system, but the regression-based scheme has a better accuracy than the one based on isotropic projection of the effective stiffness tensor. The accuracy of the projection-based scheme can be greatly improved through weighted projection.

For this 80×80×80 mesh, the DNS based on FEM typically takes 25 hours on 24 cores (in a state-of-the art high performance computing cluster with two 12-core/processor Intel Haswell E5-2680v32.5 GHz processors per compute node). With the same number of loading increments, the SCA reduced order method (in MATLAB) only takes 0.1 s, 2 s and 50 s on one Intel i7-3632 processor for km=1, 16 and 256, respectively.

Spherical Inclusion with Crystal Plasticity

The crystal plasticity law is introduced for the simplest geometric case here, that approximating the 3D Eshelby problem: a spherical inclusion/void embedded in an infinite (periodic boundary) single-crystal matrix. A schematic of the geometry is shown in FIG. 69. In the context of AM, provided in the introduction, this could be thought of as a spherical void occurring in the interior of a part. Such voids, between 1-2 microns and 50 microns diameter are often attributed to boiling and material vaporization during the build process. To model this, an nearly infinitely compressible, very low modulus material law is applied within the sphere, while the CP model is used for the matrix material. A set of crystal plasticity parameters are listed in Table 4-3; the Young's modulus and Possion's ratio of the soft inclusion are 500 MPa and 0.19, respectively. These parameters match reasonably well with a FCC metal, though are not yet calibrated for AM materials specifically.

TABLE 4-1 Crystal plasticity parameters for a FCC metal C1111, MPa C1122, MPa C2323, MPa 40356 20257   {dot over (γ)}0, s−1 m initial τ0, MPa .002   10 320  H, MPa R, MPa X   0 1 initial a0, MPa h, MPa r, MPa .0   500 0

To solve for the overall and local response of this geometry, an appropriate choice of data for the domain decomposition stage must be made. Using the strain concentration tensor and the elastic response provides a reasonable overall match in load history to the DNS solution, but the local solution (near the inclusion) does not match well. As noted above, different variables may be used to conduct the clustering. Panel (a) of FIG. 70 shows the elastic DNS solution, and panel (b) of FIG. 70 shows the clusters built from the strain concentration tensor. Choosing the plastic part of the strain tensor at the onset of plasticity, contours plotted in panel (c) of FIG. 70, results in the decomposition shown in FIG. 4d. This gives much higher cluster density near the inclusion, and we will show in our future work that this allows for much more accurate local solutions. Using the fully developed plastic solution, panel (e) of FIG. 70 gives the clustering shown in panel (d) of FIG. 70. Uniaxial tension in z direction is applied until the whole system has fully yielded. The crystal is orientated with Euler angles ψ=0°, θ=45°, ϕ=0° (using the Roe convention) with respect to a coordinate system aligned with the global axes.

Once the clusters are determined, total and local solutions for stress and strain can be computed with the SCA reduced order method with crystal plasticity, denoted as CPSCA, in the online stage. The overall solution corresponding to the clustering in FIG. 70 is shown in FIG. 71. The first set of solutions match very well in the elastic region, and begins to develop a slight difference at the onset of plasticity. CPSCA based on elasticity clustering and onset plasticity clustering gives harder response while that based on fully developed plasticity has softer response compared to CPFEM results.

For this 40×40×40 mesh, the DNS CPFEM implemented as a user material in Abaqus typically takes 4600 seconds on 24 cores. With the same number of loading increments, CPSCA (in FORTRAN) only takes 5 seconds on one Intel i7-3632 processor using 16 clusters in the matrix.

Polycrystalline RVE

In this embodiment, CPSCA is used to predict the overall response of a RVE including equiaxed, randomly oriented grains with the fully developed plastic strain tensor calculated a priori as offline database. An example of such a RVE is shown in FIG. 72. FIG. 73 shows the comparison of overall stress strain curve predicted by CPFEM and CPSCA respectively. FIG. 72 shows RVE including 35 equiaxed, randomly oriented grains (as shown by the inverse pole figure color map) with 20×20×20 and 40×40×40 voxel mesh in panel (a) and (b), respectively. FIG. 73 shows σ33 versus 33 using CPFEM and CPSCA respectively, showing convergence with mesh size and number of clusters.

We see that the overall response for the coarser cases converge to very similar solutions when element or clusters are added. CPSCA results in harder response than the CPFEM solutions when very coarse clustering (e.g., 1 cluster/grain) is used. This is not an exceptional result, because SCA uses a FFT solution based on the Lippmann-Schwinger equation.

The full 3D solution state for S33 at 5% averaged strain is shown in the opacity and color contour plots shown in FIG. 74, where panel (a) is the 203 mesh with CPFEM, panel (b) is the 303 mesh with CPFEM, panel (c) the 403 mesh with CPFEM clusters, panel (d) is the 403 mesh with 35 clusters, panel (e) is the 403 with 70 clusters, and panel (f) is the 403 mesh with 140 clusters. The 35 cluster case has one cluster per grain, whereas the 140 cluster case has four clusters per grain. Opacity scales with stress level. With this visualization, some differences in the interior can be observed: in the CPSCA method, stress is generally more concentrated, and lower outside of the concentration region, when compared to the CPFEM solutions with more distributed and generally higher levels of overall stress. In both solution methods and with all mesh sizes and number of clusters, stress concentrates in grains with high Schmid factor. The peak values for the FEM and SCA solutions are generally within 10%, while the minimum values differ by more.

Again, the DNS CPFEM implemented as a user material in Abaqus typically takes 587 seconds, 5177 seconds, and 31446 seconds for the 20×20×20, 30×30×30 and 40×40×40 mesh respectively, on 24 cores. With the same number of loading increments, CPSCA (in FORTRAN) only takes 18 seconds, 96 seconds and 793 seconds using 1 cluster/grain, 2 clusters/grain and 4 clusters/grain respectively on one Intel i7-3632 processor.

In sum, we have presented herein a method to dramatically decrease the computational cost associated with conducting microscopic crystal plasticity simulations, of the type that can be used to calibrate homogenization models, or to investigate the mechanics of processes pertaining to damage or localization within metals. This was motivated by a desire to predict the mechanical response of material resulting from the additive manufacturing process—a challenge of great interest recently. In these materials, microscale and mesoscale factors (e.g., voids of different sizes, grains sizes dependent on physical location) are of interest, necessitating a fast method to predict micromechanical solutions over a relatively large volume. The method is demonstrated with three examples: J2 plasticity in a multi-inclusion system, a simple void-like inclusion embedded within a single-crystal matrix, and a polycrystalline RVE of equiaxed grains.

Example 5 Inverse Modeling Approach for Predicting Filled Rubber Performance

In this example, a computational procedure combining experimental data and interphase inverse modeling is presented to predict filled rubber compound properties. The FFT based numerical homogenization scheme is applied on the high quality filled rubber 3D Transmission Electron Microscope (TEM) image to compute its complex shear moduli. The 3D TEM filled rubber image is then compressed into a material microstructure database using a novel Reduced Order Modeling (ROM) technique, namely Self-consistent Clustering Analysis (a two-stage offline database creation from training and learning, followed by data compression via unsupervised learning, and online prediction approach), for improved efficiency and accuracy. An inverse modeling approach is formulated for quantitively computing interphase complex shear moduli in order to understand the interphase behaviors. The two-stage SCA and the inverse modeling approach formulated a three-step prediction scheme for studying filled rubber, whose loss tangent curve can be computed in agreement with test data.

Composite materials in general exhibit improved mechanical behaviors compared to their basic constituents. Such characteristics provide a window for creating specific materials to satisfy requirements that are hard to meet by materials without specific treatments. It is well-known that the properties of Polymer Nano-Composites (PNCs) differ from pure polymer components partly due to an interphase region between polymer and filler. Thus, it is possible to design specific properties by adding fillers into polymer components (such as rubber compounds) to achieve different viscoelastic behaviors compared to pure polymers without fillers. In this exemplary study, a computational framework for the efficient evaluation of filled rubber properties and interphase property inverse modeling for improving filled rubber properties prediction is disclosed.

In the rubber and tire industry, reduction of loss tangent (or tan(δ)) can reduce rolling resistance whereas an increase of loss tangent provides more friction between the tire and the ground. Experimental studies reveal that fillers in polymer compounds indeed result in different viscoelastic behavior for PNC vs. pure polymer compound. For styrene-butadiene rubber, the addition of carbon black filler reduces tan(δ) in the low-temperature region but increases tan(δ) in the high-temperature region. Moreover, Brinson et al. conducted a study of styrene-butadiene rubber with different fillers and concluded that fillers enhanced the overall rigidity of the PNC but reduced tan(δ). Therefore, various tire properties can be achieved through custom designed polymer nano-composite (PNC), or filled rubber.

The cause of change in viscoelastic behavior between PNC and the pure polymer has been studied extensively. Due to the exponential growth of computational power over the past decades, researchers are able to utilize Molecular Dynamics (MD) simulations to capture the effect of fillers in polymer compounds by observing the interaction between polymer chains and fillers. Polymer chains aggregate around added filler material, creating a denser layer of polymers. Such results are due to van der Waals interactions between polymers and fillers. Results obtained from MD agree well with experimental observations, where the interphase exhibit stiffer responses compared to the polymer matrix. To characterize such a change of polymer structure, the interphase can be used to distinguish different viscoelastic behaviors of PNC and pure polymer compound. In this work, a filled polymer system using a three-phase model is considered in the current work: these phases are the filler, interphase, and polymer. The three-phase model is applied to the filled rubber sample studied in the present work; this model should be able to capture the difference in viscoelastic performance between filled and unfilled rubber.

Rubber properties can be experimentally measured by dynamic mechanical analysis (DMA). The experimental procedures provide viscoelastic properties, i.e., storage and loss moduli, of the rubber compound. As a result, DMA provides overall homogenized properties of the rubber. Numerically, macroscopic properties can be determined through various homogenization approaches. The aforementioned 3D TEM process has been applied to a different filled rubber geometry which has already been studied via the Finite Element (FE) method to obtain the local stress response. However, in this past work, the filled rubber geometry provided by 3D TEM as a 3D digital image was only converted to a fine conforming FE mesh which could only be used on a supercomputer. Voxel meshes of large sizes such as 3D digital images provided by 3D TEM are more suitable for computational homogenization using the FFT based numerical method. The digital image can be directly used in the FFT based algorithm to solve for local and overall responses under designated macroscopic boundary conditions. Therefore, the properties of the filled rubber can be directly obtained via the FFT algorithm from using the 3D image of the filled rubber.

The existing experimental techniques allow measurement of tan(δ) curves (also known as master curves), of filled and unfilled rubber through DMA. Therefore, if the filled rubber is assumed to be a two-phase model with rubber and filler phases, it is possible to conduct a numerical simulation of its responses at different frequencies using measured properties of unfilled rubber and filler. In the present work, the property of interphase is unknown, which makes it hard to predict the filled rubber's master curve through basic constituents: unfilled rubber and filler. In this exemplary example, we combined interphase modeling via the so-called inverse modeling technique and an efficient FFT-based homogenization scheme to compute the master curve of a filled rubber using a fine mesh reconstructed from 3D TEM. The master curve is validated by experimental results to illustrate the efficacy of our proposed scheme. Comparison of the FFT method and the SCA is performed to examine the improved computational efficiency of this reduced order modeling-based approach.

The following sections focus on the experimental study of filled rubber and 3D TEM reconstruction of a filled rubber sample, details on the FFT algorithm employed, the SCA method, the inverse modeling scheme for interphase properties computation, and results and discussion. We demonstrate an efficient integrated experimental and modeling approach that combines the merits of an FFT homogenization algorithm, data-driven SCA, with an inverse modeling technique. With 3D TEM reconstruction of a filled rubber sample and experimental DMA data, we gain an understanding of interphase properties in filled rubber components.

Experimental Study of the Filled Rubber

Preparation of Unfilled and Filled Rubber Samples

The unfilled rubber sample used in the present study is made of poly-isoprene (IR2200) with some chemical agents such as sulfur, stearic acid, microcrystalline wax, zinc oxide and etc. In the first stage of mixing, poly-isoprene and agents other than curing agents were mixed. Then curing agents, such as sulfur, were added. The mixture of polymer and agents was cured with a high temperature to obtain a vulcanized unfilled rubber sample. The filled rubber sample is made of poly-isoprene, silica particles (Zeosil 1165MP), and some chemical agents. It was obtained in the same way except that silica particles are added in the first stage of mixing. The formula of these rubber samples is summarized in Table 5-1.

TABLE 6-1 Formula of rubber samples (weight ratio per hundred rubber) Unfilled rubber Filled rubber sample sample Isoprene (IR2200) 100 100 Sulfur (cross-linking agent) 1.5 1.5 Other agents 9.3 9.3 Silica (Zeosil 1165MP) None 40

Experimental Results on Master Curve Measurements of Filled and Unfilled Rubber Samples

Storage and Loss modulus of filled and unfilled rubber samples were measured by a dynamic mechanical analysis (DMA) (TA Instruments, rubber rheometer ARES-G2). Cylindrical samples with a diameter of 8 mm and a height of 6 mm were used. The measurements were operated at 0.1% oscillatory shear strain in which the material response is in the linear viscoelasticity region. In order to obtain the viscoelastic responses in a wide frequency range, a frequency sweep from 0.5 Hz to 50 Hz was operated at a temperature range of −60° C. to 40° C. Master curves were obtained by time-temperature superposition with the reference temperature of 25° C. where only horizontal shift was performed. It is clearly indicated in FIG. 75 that the temperature dependence of the horizontal shift factor aT is well described by the Williams-Landel-Ferry (WLF) equation. FIG. 76, FIG. 77, and FIG. 78 show the storage shear modulus, loss shear modulus, and loss tangent of these samples respectively.

3D-TEM Reconstruction of Filled Rubber Sample

The filled rubber was made into thin sections with about 100 nm thickness using a focused ion beam (JEOL, JEM-9310FIB) at a cryogenic temperature at −150° C. The ultrathin section was transferred onto a Cu mesh grid with a polyvinyl formal supporting membrane. Prior to the electron microscopy experiments, gold particles 5 nm in diameter were placed on the ultrathin section from colloidal aqueous solution.

We conducted 3D observations by TEM and 3D-TEM using a JEM-2200FS microscope (JEOL, Ltd., Japan) operated at 200 kV. The electron microscope was equipped with a slow-scan USC 4000 CCD camera (Gatan, Inc., USA). Elastically scattered electrons (electron energy loss: 0±40 eV) were selected by an energy filter installed in the microscope (Ω filter, JEOL Ltd., Japan).

A series of TEM images were acquired at tilt angles in the range of −66° to 73° at an angular interval of 1°. Subsequently, the tilt series of the TEM images were aligned by the fiducial marker method, using gold nanoparticles as the fiducial markers. The tilt series of TEM images after the alignment were reconstructed by filtered back-projection (FBP). It took us about 2 hours and a few days, respectively, to take 140 TEM tilt images on TEM and to align those projections before the FBP reconstruction. In addition, segmenting out the fillers from the rubber matrix in each digitally-sliced image has been done before stacking them to generate a 3D image, which takes, typically, 1 to 2 weeks. The basic protocol used here is essentially the same as the one in which we demonstrated less than 1 nm resolution, i.e., 0.5 to 0.8 nm. The reconstructed filled rubber is shown in FIG. 79, where 3D-image of measured filler structure is shown in top view and in front view, and rubber matrix material is hidden.

Fast Fourier Transform Homogenization Scheme

Formulation of Fast Fourier Transform Scheme

3D TEM generates a high-fidelity 3D reconstruction of a filled rubber sample as a 3D digital image with a resolution of the sub-nanometer, as mentioned in the previous section. The nature of 3D TEM leads to very large structured voxel meshes making it hardly feasible for FE computation due to the fine resolution. The FE mesh was generated directly from the high-resolution 3D TEM data as a voxel mesh with a large total number of degrees of freedom. To make the reconstruction suitable for FE analysis, the 3D digital image had to be converted into a conforming mesh. However, such process required extra time and resources in preprocessing before the actual simulation. Recent development in FFT homogenization scheme provides an alternate solution for solving boundary value problems using a structured mesh. The FFT homogenization scheme avoids this conversion process and makes computation straightforward using the 3D digital image from the 3D TEM process. Although the convergence of the FFT scheme for arbitrary phase contrasts and its efficiency can still be improved, it is a powerful homogenization scheme for 3D images. For example, the FFT scheme bypasses the need for mesh generation, which is required by the FE method, and reduces the problem size so that it can be solved on a personal computer. Although the large input image might require larger computer memory due to an increase of the total number of degrees of freedom, the FFT scheme enables a much easier approach when a high-resolution 3D digital image is provided.

In this exemplary example, we first adopt the FFT scheme assuming small strain based on the system of equations shown in Eq. (5-1).

{ . σ ( X ) = 0 , X Ω 0 ( equilibrium ) σ ( X ) = f ( X , ɛ ) ( constitutive law ) ɛ = s y m ( u * ) + ɛ M a c r o ( compatibility ) ( 5 - 1 )

where the applied displacement field u* is periodic over the computation domain Ω0. The equilibrium condition for any input mesh is given as below following conservation of linear momentum:


∇·σ(X)=0,∀X∈Ω0  (5-2)

In the FFT scheme, each voxel in the input image represents a material point. The location of the voxel is represented by X. Stress and strain tensors at all material points are computed in the FFT scheme. The local stress σ(X) can be computed using any given constitutive law, but we assume linear elasticity for now.

To solve Eq. (5-2), the FE approach would formulate the Cauchy momentum equation with periodic boundary conditions. Here, the local strain is given as in Eq. (5-1). It is composed of sym(∇u*), the symmetric part of the gradient of the periodic displacement field u*, and εMacro, the prescribed macroscopic strain tensor. Introducing the polarization stress T and the Green's operator Γ0, it is possible to express to the local strain as a Lippmann-Schwinger equation:


ε(X)=−(0*τ)(X)+εMacro  (5-3)

The polarization stress τ and the explicit form of the Green's operator 0 in Fourier space are defined in Eq. (5-4)


τ(X)=0:ε(X)−σ(X)  (5-4)

where 0 is the standard stiffness tensor of an isotropic reference material, written as Cklmn00δklδmn0 δkmδln in index notation, with reference Lamé parameters λ0 and μ0.

Γ ^ klmn 0 ( ξ ) = δ k m ξ l ξ n 2 μ 0 ξ 2 - λ 0 2 μ 0 ( λ 0 + 2 μ 0 ) ξ k ξ l ξ m ξ n ξ 4 ( 5 - 5 )

where the indices of {circumflex over (Γ)}klmn0 coincide with those of Cklmn0.

Since the explicit form of the Green's operator is known only in Fourier space, the convolution term in Eq. (5-4) is computed with the help of the inverse Fourier transform as in Eq. (5-6).


0*τ(X)=−1{{circumflex over (Γ)}klmn0(ξ)[τkl(X)]}  (5-6)

where and denote respectively the FFT and the inverse FFT.

To solve the above FFT formulation, various iterative methods can be used, such as fixed point iteration and conjugate gradient. The solution techniques are vastly available in the literature. For demonstration purposes, the FFT algorithm is presented using fixed point iteration in Appendix A.

The iteration process of the FFT algorithm starts from a given initial local strain and checks for convergence on the local ε(X). During the iteration process, Green's function enforces the compatibility condition given in Eq. (5-1). To obtain the macroscopic stress and strain tensors from the mesh, volumetric average following Hill's lemma are conducted as in Eq. (5-7):

σ Macro = 1 V V σ ( X ) dV and ɛ Macro = 1 V V ɛ ( X ) d V ( 5 - 7 )

To obtain effective elastic material properties μMacro and λMacro, σMacro and εMacro are plugged into Hooke's law as in Eq. (5-8):


σijMacroMacroεkkMacroδij+2μMacroεijMacro  (5-8)

It is convenient to compute μMacro and λMacro by solving Eq. (5-8). One can re-write Eq. (5-8) in matrix format and solve for μMacro and λMacro.

Application of FFT Scheme for Frequency Domain Computation

For rubber materials or viscoelastic materials in general, responses are drastically different for various loading frequencies; DMA is a common experimental method to evaluate this variation. DMA provides viscoelastic material properties, such as the complex Young's modulus, E*=E′+iE″, and the complex shear modulus, at different frequency points, denoted as ωk. The ratio between the E″, the imaginary part of the complex Young's modulus, and E′, the real part of the complex Young's modulus, gives the tan(δ) curve. Complex young's modulus can be interchangeable with complex shear modulus G*=G′+iG″.

For DMA, a sinusoidal strain with a given frequency ωk is applied to the rubber and the steady state stress is measured to compute viscoelastic material properties. In the 1D case, the stress at a given peak strain ε0Macro can be written as:

σ M a c r o ( t ) = [ ɛ 0 M a c r o E * ( ω k ) e i ω k t ] = [ ɛ 0 M a c r o ( E ( ω k ) + iE ( ω k ) ) ( cos ( ω k t ) + i sin ( ω k t ) ) ] ( 5 - 9 )

where, for any complex number z, [z] is the real part of z.

A steady-state solution for σMacro(t) can be found by simply taking

t = 0 , 2 π ω k , 4 π ω k ,

The steady state stress can then be treated as a complex one, written as σMacro,*0Macro(E′(ωk)+iE″(ωk). σMacro,* is composed of a real part and an imaginary part. For the 1D case, taking the quotient of σMacro,* and ε0Macro yields the complex Young's modulus. Reciprocally, the complex stress can be computed by the FFT scheme by inputting the complex Young's modulus.

The above operation is to be performed at a given frequency point ωk. To compute the rubber's tan(δ) curve, an individual tan(δ) point at different ωk is needed, where tan(δ) is defined as

tan ( δ ) = E E

for tensile DMA and

tan ( δ ) = G G

for shear DMA. Therefore, by computing the rubber's responses at different frequency points, the complex Young's modulus or shear modulus can be obtained and tan(δ) can then be computed. When a sufficient number ωk is taken, a smooth tan(δ) curve of filled rubber can be reconstructed.

In this exemplary example, shear DMA of both unfilled and filled rubber are conducted to reconstruct master curves. However, due to the limitation of experimental conditions, only complex shear moduli at different frequencies are available. An assumption is that for a complex shear modulus, a conversion to the Lamé constants is still valid. This enables computation of tan(δ) using the FFT scheme at various frequency points for the filled rubber by inputting properties of basic constituents: unfilled rubber and fillers.

A computation using unfilled rubber complex shear moduli at different frequency points and filler properties is performed using the mesh introduced above. To make the computation more feasible, the filled rubber domain is shrunk by ½ in all three directions to 513×513×75 voxels, where the length of each voxel's edge is 1.62 nm. Poisson's ratio of the unfilled rubber is taken as 0.499 allowing for limited compressibility of rubber materials. Note that the aforementioned assumption can be improved by choosing a frequency dependent Poisson's ratio to account for limited compressibility in the glassy state, given enough experimental data from material characterizations. This will be addressed in future work. The filler material has Young's modulus E=300 MPa and Poisson's ratio ν=0.19. The homogenized complex shear modulus from the FFT scheme is used to compute the tan(δ) curve. The computed tan(δ) curve and the experimental result are shown in FIG. 80.

The results in FIG. 80 illustrate the inconsistency between prediction and experimental measurement of tan(δ). It can be deduced from the figure that before the peak of tan(δ), the two-phase model follows a different trend than experimental results. In the low-frequency region (less than 1e5 Hz), tan(δ) is lower than experimental measurements. Between 1e5 Hz and 5e7 Hz, tan(δ) is overpredicted.

The predicted G′ and G″ of filled rubber model are shown in panels (a)-(b) of FIG. 81. In panel (a) of FIG. 81, the prediction of G′ in the low-frequency region, i.e., between 1e1 Hz and 1e5 Hz, is 50% smaller than experimental values. In the same region, shown in panel (b) of FIG. 81, be predicted loss modulus G″ can be as low as 100% below the experimental values. Such discrepancy between simulation and prediction result in lower tan(δ) as shown in FIG. 80. In the high-frequency region, between 1e5 Hz to 1e7 Hz, the predicted storage modulus has the same 50% difference w.r.t. experimental data, but the difference between the predicted G″ and experimental data decreases, resulting in a trend of increasing G″. This explains why the predicted tan(δ) shown in FIG. 80 between 1e5 Hz and 1e7 Hz is higher than the experimental data. In the region of frequency higher than 1e7 Hz, both predicted G′ and G″ are close to experimental results, giving a relatively good prediction of tan(δ).

Such an inconsistency observed in FIG. 80 is expected, as the input mesh only considered rubber and filler phases. As mentioned in the introduction, the existing literature supports the assumption that there is an interphase region between the filler and pure rubber that should be treated as a third material. Specifically, the interphase should have a larger G″ compared to unfilled rubber in the lower frequency region between 1e1 Hz and 1e5 Hz. It should also have a larger G′ compared to the unfilled rubber in the frequency region between 1e5 Hz and 1e7 Hz. The original two-phase model is incomplete and cannot fully reveal the property of filled rubber, which would be required to model the interphase based on the filled rubber 3D image reconstructed. One problem that emerges after modeling the interphase in the original filled rubber mesh is that the property of the interphase stays unknown. Due to a limitation of experimental techniques, the viscoelastic property of interphase, or its complex Young's or shear modulus cannot be given as a known input. Therefore, we propose an inverse modeling technique to compute the interphase property quantitatively, described in the next section, with the aim of producing a better prediction of tan(δ) of filled rubber. However, the inverse modeling procedure requires multiple iterations and FFT procedure is performed in all iterations. This will impose a considerable computational cost. To address the computational efficiency issue, a reduced order modeling approach is also introduced to reduce the computational cost of evaluating filled rubber responses at each frequency point. The reduced order modeling approach is then combined with the inverse modeling procedure to compute interphase properties and filled rubber properties.

In terms of the effect of filler moduli, Appendix C gives predicted G′ and G″ of the same filled rubber structure with filler Young's moduli from 300 MPa (3E+8 Pa) to 3 GPa (3E+9 Pa). The variation of the filled rubber G′ and G″ is very little. This means the predicted filled rubber is insensitive to the filler modulus. However, since the FFT algorithm is sensitive to the contrast between filler and matrix materials, filler Young's modulus of 300 MPa is used in the present work.

Efficient Reduced Order Modeling for the Filled Rubber Composite

The aforementioned FFT formulation can compute the effective property of filled rubber, but it requires the computation of all local responses at individual voxels and thus imposes a high computational cost, though much less when compared to the FEM. The recently proposed SCA method provides an alternative for computing effective properties of arbitrary microstructure, such as the filled rubber composite, at reasonable a computational cost. In this section, the SCA formulation is discussed, providing insight into the physically-based reduced order model.

SCA is a two-stage reduced-order modeling approach. In the offline stage, two steps are performed: 1) all voxel elements in the mesh are clustered based on arbitrary measurement of similarity in mechanical responses, such as strain concentration tensor ; 2) The interaction tensor, , for each pair of clusters is then computed. The offline stage will generate a material microstructure database which contains all interaction tensors between clusters pairs and volume fraction of each cluster. After the offline state, the original high fidelity RVE is compressed into a small number of clusters. In the online stage, discretized Lippmann-Schwinger will compute strain and complex stress in each clusters and RVE level averaged complex stress and strain at any given external loading conditions by solving a boundary value problem. Once the RVE complex stress is identified, tan(δ) will be computed accordingly. This process is concisely illustrated in FIG. 82. Note that in the previous section, RVE complex stress at each frequency point is computed by FFT. Here, voxels are assumed to be “once responded same, always responds the same”. Therefore, the offline database can be constructed by once. The database can be used for all frequency points afterward for an efficient evaluation of tan(δ) at different frequency points.

The Lippmann-Schwinger equation given in Eq. (5-3) can be reformulated in the following form in Eq. (5-10)


εMacro−ε(X)−∫Ω0(X,X′):[σ(X′)−0:ε(X′)]dX′=0,X∈Ω  (5-10)

where X is the voxel element location in the mesh and X′ is the location in the reference domain.

To perform reduced order modeling of the filled rubber composite, one can reduce overall degrees of freedom in the mesh by grouping voxels with similar mechanical responses together. This process is also known as clustering. A convenient criterion chosen here is the well-known strain concentration tensor that connects macroscopic prescribed strain to local strain responses at each voxel, shown in Eq. (5-11) below:


ε(X)=(X):εMacro,X∈Ω  (5-11)

where is a 6 by 6 matrix in Voigt notation. Six traction free loadings on the original RVE are needed in order to determine all 36 entries of . Clustering algorithms, such as k-means clustering, can be used to cluster all voxels and decompose the original domain of 19,737,675 voxels into K clusters, where K equals 64. This is the first step of the offline stage. One might think this as reducing total integration points to K, where K is a smaller number comparing to the total number of integration points in the original mesh. Note that is not the sole solution for the clustering process. For different problems, one might wish to use other meaningful quantities to identity voxels with similar mechanical responses, such as lattice orientation.

It is convenient to define a characteristic function as in Eq. (5-12) in order to decompose Eq. (5-10) to incorporate the newly decomposed domain.

χ I ( X ) = { 1 X Ω I 0 otherwise ( 5 - 12 )

where I=1,2,3, ,K. The discretized Lippmann-Schwinger equation is given in Eq. (13) for each cluster.

ɛ M a c r o - 1 c I Ω Ω χ I ( X ) ɛ ( X ) dX - 1 c I Ω Ω Ω χ I ( X ) Γ 0 ( X , X ) : [ σ ( X ) - 0 : ɛ ( X ) ] dX dX = 0 , X Ω ( 5 - 13 )

where cI is the volume fraction of cluster I and |Ω| is the total volume of the mesh.

By noticing that σ(X′) and ε(X′) can be written as:

σ ( X ) = J = 1 K χ I ( X ) σ J , ɛ ( X ) = J = 1 K χ J ( X ) ɛ J ( 5 - 14 ) Now Eq . ( 13 ) becomes : ɛ M a c r o - 1 c I Ω Ω χ I ( X ) ɛ ( X ) dX - 1 c I Ω J = 1 K Ω Ω χ I ( X ) χ J ( X ) Γ 0 ( X , X ) : [ σ ( X ) - 0 : ɛ ( X ) ] dX dX = 0 , X Ω ( 5 - 15 ) where 𝔻 IJ is : 𝔻 IJ = 1 c I Ω Ω Ω χ I ( X ) χ J ( X ) Γ 0 ( X , X ) , X Ω ( 5 - 16 )

After the first step of the offline stage process, which is the clustering process, IJ can be computed. Once IJ is computed, the second step of the clustering process is completed and the original RVE is compressed into a microstructural database made of clusters and interaction tensors. Plug in IJ into Eq. (5-13) will give the final form of the discretized Lippmann-Schwinger equation in Eq. (5-17):

ɛ M a c r o - ɛ I - J = 1 K 𝔻 IJ : [ σ J - 0 : ɛ J ] = 0 , I = 1 , 2 , 3 , , K ( 5 - 17 )

where the incremental form is given as:

Δ ɛ M a c r o - Δ ɛ I - J = 1 K 𝔻 IJ : [ Δ σ J - 0 : Δ ɛ J ] = 0 , I = 1 , 2 , 3 , , K ( 5 - 18 )

The online stage involves the evaluation process of Eq. (5-18). The solution procedure of Eq. (5-18) is given in Appendix B for readers' reference.

The above SCA formulation is combined with Eq. (5-9) to compute the effective complex moduli of filled rubber at a reduced computational cost due to the reduction of the number of integration points and degrees of freedom. The ROM of the original filled rubber domain with clusters is shown in FIG. 83 for 32 clusters per phase. A second ROM with 64 clusters per phase (making 128 clusters in total) is also generated but not shown for the sake of space. The corresponding master curve is given in FIG. 84, where the comparison with the FFT result is shown. The master curves computed by the ROM show a trend of the overall effective filled rubber performance that is consistent with the results computed by the FFT algorithm. Although SCA over-predicts tan(δ) curve, the results were obtained with a dramatic reduction of computation time with a speedup of 924 at 64 clusters (32 clusters in the filler phase and 32 clusters in the unfilled rubber phase). The comparison between computation time using SCA and FFT are shown in Table. 2 below. In FIG. 85, the filled rubber G′ and G″ computed by SCA are in the same trend. However, SCA provides considerable savings in terms of computation time. The difference between FFT and SCA predictions are within one order of magnitude, this means the ROM doesn't sacrifice all microstructure information after the clustering process. Both FFT and SCA predict the same trend of filled rubber G′ and G″, meaning that the deviation caused by the interphase needs to be captured as mentioned in the previous section.

To further investigate interphase properties using our reduced order model approach, the above procedure is integrated into the inverse modeling process to better predict filled rubber properties.

TABLE 5-2 Comparison of Computation Time over 17 Frequency Points Computation Time FFT 4023 s per frequency point, 68400 s total CPU time ROM (64 clusters) Offline: 7 hr (strain concentration tensor generation) 5 hr (clustering + interaction tensor generation) Online: 4.35 s per frequency point, 74 s total CPU time

Inverse Modeling Scheme for the Interphase

With the master curves shown in FIG. 84, it is clear that a two-phase filled rubber model is not sufficient in representing the true filled rubber behavior. The cause of this inconsistency between computed tan(δ) curves and experimental measurement can be attributed to the interphase region between the filler material and unfilled rubber. Therefore, in the following discussion, the presence of the interphase between filler and matrix materials is considered. This should reduce the difference between the tan(δ) curves plotted in FIG. 84. The interphase is assumed to surround the filler in a spherical region with a radius βIP of 9.74 nm. Note here that the interphase is assumed to be uniform, meaning its thickness is constant throughout the whole domain. A voxel-wise search of the filled rubber 3D-image is performed to convert the elements within βIP from the matrix to the interphase.

Inverse Modeling Formulation

The filled rubber model with the interphase can be created following the aforementioned voxel-wise search process, but the viscoelastic behavior of interphase is still unknown. The interphase is used to suppress the inconsistency between master curves from the FFT homogenization scheme and experimental data, thus its complex Young's or shear moduli at different ωk have to be computed. To predict the unknown interphase property with limited experimental data, a so-called inverse modeling scheme is introduced based on optimization techniques. The objective function of the inverse modeling process can be written as:

G * , I P ( ω k ) = ( ω k ) = min ( ω k ) + ( ω k ) norm [ ( G , PMC ( ( ω k ) , ω k ) - G , EXP ( ω k ) ) + i ( G , PMC ( ( ω k ) , ω k ) - G , EXP ( ω k ) ) ] ( 5 - 19 )

Above goal function states for each given ωk, the solution of interphase complex shear modulus G*′IPk) is found when the difference of predicted complex shear modulus G*′PMC(G*′IP, ωk) and G*′EXPk) is minimized. When ωk is fixed, it is possible to define a function for the root-finding process as:


f(G*′IP)=G*′PMC(G*′IP)−G*′EXP  (5-20)

where ωk is omitted compared to Eq. (5-10) since the solution is found for each ωk of interest.

To solve for Eq. (5-20), an iterative method is used to find the solution of G*′IPk). The derivative of Eq. (5-20) can be formulated as in Eq. (5-21) in order to apply Newton's iterative method:

f ( G * , IP ) = ( G * , PMC , n - G * , E X P ) - ( G * P M C n - 1 - G * , EXP ) G * , IP n - G * , IP n - 1 ( 5 - 21 )

With Eq. (5-20) and Eq. (5-21), it is possible to write the iterative process as Eq. (5-22):

G * , IP n + 1 = G * , IP n - G * , PMC n - G * , EXP [ ( G * , PMC n - G * , EXP ) - ( G * , PMC n - 1 - G * , EXP ) G * , IP n - G * , IP n - 1 ] ( 5 - 22 )

where the superscript n denotes the current iteration number and n+1 denotes the next iteration. The initial guess for interphase properties is set to be the same as an unfilled rubber.

Inverse Modeling for the Filled Rubber Model with Interphase

The proposed inverse modeling method is applied to the aforementioned filled rubber domain to re-compute the filled rubber master curve. The mesh is the one used above, but with the added interphase. Interphase of thickness βIP=9.74 nm, which is equivalent to 6 voxels, is added to the domain to create the filled rubber model with the interphase. The updated filled rubber mesh with the interphase is shown on the left of FIG. 86 below. The corresponding ROM of the above three-phase model has also been constructed and shown on the right of FIG. 167. FIG. 86 shows the ROM with 32 cluster in each phase, meaning the whole filled rubber mesh is compressed into 96 clusters.

For the inverse modeling process, 17 frequency points are picked over the span of the entire master curve of filled rubber. More points can be used for the inverse modeling process but would increase computational cost. The experimental procedure of measuring the filled and unfilled rubber master curves have been reported above. Material properties for unfilled rubber and filler materials are also introduced above. The inverse modeling process is combined with the SCA online prediction to be the third step of the present scheme. At this point, the three-step prediction scheme for filled rubber is presented. The results of the inverse modeling will be presented and discussed in the next section.

Result and Discussion on Inverse Modeling Results

Through inverse modeling, a more accurate prediction of the filled rubber master curve is shown in FIG. 87. Predicted tan(δ) curves by both the FFT and SCA method are in a good match with measured experimental data. The SCA result is consistent with the FFT result, meaning the current ROM provides sufficient accuracy in predicting overall behavior of the filled rubber.

The comparison between the predicted G′ and G″ of the filled rubber and experimental results are shown in FIG. 88 a) and b), respectively. It can be concluded that with the consideration of the interphase, the prediction of tan(δ) reaches a good match with the experimental data.

Through inverse modeling, G′ and G″ of interphase are computed as well, as shown in panels (a)-(b) of FIG. 88. As stated above, the interphase should have larger G′ and larger G″ compared to unfilled rubber properties. Only then the filled rubber G′ and G″ can be improved and match with experimental results. The inverse modeled interphase properties ensure the prediction of the filled rubber's behavior follows the experimental data. It is worth noting the interphase G′ and G″ predicted by the ROM through SCA is higher than that obtained from FFT, but both SCA and FFT predict the same trend of the filled rubber G′ and G″ as a function of frequency. Such a deviation is expected since the G′ and G″ predicted using a 2-phase filled rubber by SCA are smaller than predictions made by FFT. Higher interphase G′ and G″ will offset such differences in order to comply with the objective function defined in Eq. (5-19).

The difference observed between the filled rubber and unfilled rubber tan(δ) shown in FIG. 87 suggested that the filled rubber would have less rolling resistance due to smaller tan(δ) measured at a lower frequency range (1e4 Hz and 1e6 Hz). It is possible to link such characteristics to computed interphase properties, in which higher storage and loss moduli are observed, shown in FIG. 88. This leads to a lower tan(δ) of the interphase, leading to a decreased overall hysteresis of the filled rubber by reducing filled rubber tan(δ). Hence, the filled rubber would have less rolling resistance compared to the pure rubber.

On the other hand, FIG. 87 shows little difference between the filled and unfilled rubber tan(δ) in the high-frequency range. This suggests that in the high-frequency range both materials should behave in a similar fashion. The computed interphase storage and loss moduli do not vary much from the unfilled rubber properties as shown above. Therefore, the computed interphase properties are consistent with measured tan(δ) of the filled rubber compound.

Also, the peak tan(δ) of the filled rubber is lower than unfilled rubber. In order to increase tire traction at low temperatures, tan(δ) at lower temperatures should be increased. Based on the computed filled rubber properties, this can be achieved by identifying filler material that can form interphase with high loss modulus in the high-frequency range. Such a material combination would provide increased damping so the winter traction can be improved. The present workflow is suitable to inversely model necessary interphase storage and loss moduli that narrow down the domain for material selection.

In terms of inverse modeling results obtained from SCA, FIG. 87 shows higher interphase G′ and G″ magnitudes by SCA prediction. This is due to the lower filled rubber G′ and G″ prediction made by SCA as shown in FIG. 85 and the role of the interphase to compensate the difference observed between the 2-phase filled rubber structure and the 3-phase filled rubber structure. Hence, larger magnitudes of interphase G′ and G″ means a stiffer mechanical response of filled rubber, which explains lower tan(δ) observed in FIG. 78. In terms of viscoelastic behavior, both SCA and FFT can capture the glass transition region, as described by the peak of tan(δ) shown in FIG. 87.

In addition, a dramatic increase in computational efficiency was observed for the SCA prediction. The comparison of the computation time of inverse modeling at a single frequency point is shown in the Table 5-3 below, where a 1778 speed-up is achieved by applying the proposed ROM compared to FFT. A comparison of computational time in evaluating filled rubber properties by different methods is summarized in Table 5-4. Even though SCA requires an offline stage computation to generate the ROM of the filled rubber composite, which is still computationally expensive. However, once the database is computed, it can be used for all later evaluation procedure. This provides considerable savings in computational time in the online stage prediction of effective properties of the filled rubber, as well as the inverse modeling process. The SCA method, combined with inverse modeling, opens a new avenue towards material design. Moreover, the same sets of ROM can be used for various material properties to compute filled rubber mechanical behaviors at a reasonable cost. It is possible to explore the design space and get both a decent trend and quantitative description of filled rubber. More importantly, the proposed scheme provides an efficient solution towards investigating interphase properties of filled rubber materials for future design needs even with limited computational resource.

TABLE 5-3 Comparison of Inverse Modeling CPU Time at Single Frequency Point Computational CPU Time FFT about 35,576 s ROM (96 20 s clusters)

TABLE 5-4 Speed Comparison of Filled Rubber Property Evaluation for a model of size 513 × 513 × 75 using Different Methods Degrees of Freedom Memory Method (count) Required Computation time FEM 200 million 1 TB 1 million hr. CPU time (with 6144 cores on a supercomputer) FFT 118 million 53 GB 9.88 hr. per frequency point, 168 hr CPU time (5,952 speed-up) SCA (96 Offline: Offline: Offline: 19 hr clusters) 118 million 53 GB 11 hr (strain concentration tensor Online: Online: generation) 576 less than 8 hr (clustering + interaction 1 GB tensor generation) Online Prediction: 20 s per frequency point, 340 s CPU time (52,631 speed-up)

Despite the encouraging results observed using the inverse modeling scheme, necessary assumptions were made for this process to be possible. For example, the interphase thickness βIP was assumed to be in a circular region around the filler material. However, it is possible for the interphase thickness to be a function of the filler curvature or filler size since the degree of polymer chain aggregation can be affected by such parameters. The scheme will be extended to include varying interphase thickness around filler to consider geometrical effect.

In this exemplary study, an inverse modeling scheme is introduced and illustrated as an effort of quantitatively analyzing the interphase properties of a filled rubber compound using high fidelity reconstruction of the filled rubber sample. The FFT scheme enables efficient computation when the fine 3D digital image is used as input. The test data of unfilled and filled rubber provide enough inputs to solve an inverse modeling process for interphase properties at each frequency point. In addition, SCA, a reduced order modeling scheme, is combined with the inverse modeling procedure to compute interphase properties for the first time. Once the offline stage database is constructed, the database can be conveniently used at all frequency points to compute the whole filled rubber master curve. This novel reduced order modeling approach provides considerable savings in the computational cost. The consolidation of SCA and the inverse modeling scheme is an efficient and valuable filled rubber design tool. The present method is general enough and can incorporate other details of the microstructure, such as variation of interphase thickness. The obtained interphase property can enable forward computation of a three-phase filled rubber model in the time domain analysis, such as tensile testing. It is believed that the effect of interphase can yield better predictions of filled rubber responses under various loading conditions, and it shall be addressed in future work.

FFT Scheme Algorithm Flow Chart

The algorithm flow chart for the FFT scheme with fixed point iteration is concisely given as below. The convergence test is used to determine if the local strain εi+1 reached a stable value or not. The implementation can be easily done in any programming language, provided that FFT and inverse FFT packages are readily available.

Initialization:


ε0(X)=εMacro∀X∈V;


σ0(X)=(X):ε0(X),∀X∈V;

Iterate i+1 with εi and σi known;

    • a)={circumflex over (σ)}ii);
    • b) Convergence test;

{ ɛ ^ i + 1 ( ξ ) = ɛ ^ i ( ξ ) - Γ ^ 0 ( ξ ) : σ ^ i , ξ 0 ; ɛ ^ i + 1 ( 0 ) = ɛ M a c r o c ) ɛ i + 1 = - 1 ( ɛ ^ i + 1 ) ; d ) σ i + 1 ( X ) = ( X ) : ɛ i + 1 ( x ) ; e )

Above algorithm flow chart is for single loading step. Readers can easily modify it to multiple loading steps by defining multiple εMacro for multiple loading steps.

Self-Consistent Clustering Online Analysis Solution Procedure

SCA requires solving the discretized Lipmann-Schwinger equation based on external loading condition, either the fixed strain increment ΔεMacro or the fixed stress increment ΔεMacro. The discretized Lipmann-Schwinger equation is shown as in

Δ ɛ M a c r o - Δ ɛ I - J = 1 K 𝔻 IJ : [ Δ σ J - 0 : Δ ɛ J ] = 0 , I = 1 , 2 , 3 , , K ( 5 - 23 )

The solution to Eq. (5-23) would be strain tensor εI in each cluster. In order to use Newton's Raphson method to find a solution to Eq. (5-23), the residual form is given as in Eq. (5-24) below

r I = - Δ ɛ M a c r o + Δ ɛ I + J = 1 K 𝔻 IJ : [ Δ σ J - 0 : Δ ɛ J ] , I = 1 , 2 , 3 , , K ( 5 - 24 )

For macro strain boundary condition, the residual of macroscopic strain is written as

r K + 1 = - Δ ɛ M a c r o + I = 1 K c I Δ ɛ I , I = 1 , 2 , 3 , , K ( 5 - 25 )

For macro stress boundary condition, the residual becomes

r K + 1 = - Δ σ M a c r o + I = 1 K c I Δ σ I , I = 1 , 2 , 3 , , K ( 5 - 26 )

Solving for ΔεI by minimizing residual rI. Linearizing rI with respect to Δεyields

0 = r I + r I Δ ɛ I δ ɛ I , I = J = 1 , 2 , 3 , , k , k + 1 ( 5 - 27 )

where Jacobian Matrix

𝕄 IJ = r I Δ ɛ J

is


IJIJ+:(−0),I=J=1,2,3, . . . ,k  (5-28)

For macroscopic strain boundary condition, one has:


I(k+1)=−,(k+1)I=cI, and (k+1)(k+1)=0,I=1,2,3, . . . ,k  (5-29)

For macroscopic stress boundary condition, one has:


I(k+1)=−,(k+1)I=cI, and (k+1)(k+1)=0,I=1,2,3, . . . ,k  (5-30)

Solving Eq. (5-27) gives 6c1 that updates all local strain increment ΔεI. This process should be repeated until residuals in all clusters are minimized.

Example 6 Image-Based Multiscale Modeling System for Mechanical Performance of Metal Additive Manufacturing

The common assumptions and methods used for multiscale or microstructure-sensitive modeling of materials are generally not appropriate for capturing the performance of additively manufactured (AM) metal. Current approaches often rely upon a RVE or some other form of representative structure (e.g., a representative unit cell or simple periodic structure); prediction of response with these structures might also be predicated upon simplifying assumptions, such as idealization of the microstructure (e.g., as an ellipsoid), periodicity, or statistical (spatial) uniformity. Models that predict minimum performance or capture worst-case scenarios often struggle to make predictions that are quantitatively useful in process design—one of the main goals of structure-properties-performance modeling. These difficulties relate to an underlying challenge with AM: the localized process produces a mix of process-dependent and random microstructures, which do not follow either roughly deterministic (where worse-case analysis might be useful) or fully stochastic (where statistical uniformity would apply) patterns. The material that results from metal AM is heterogeneous, non-uniform, and highly variable.

In this example, a new paradigm is introduced in mechanical modeling of microstructure dependent failure wherein a fast reduced order approach is applied directly to experimentally imaged microstructures which populate a macroscale component; the behavior of these microstructures is used to predict the macroscale performance. Knowledge of the process history (either from modeling or experiments) can be used to select which microstrcture occurs a each material point in the component-level model. Thus spatial, build-to-build, and part-to-part variations can be captured. Specifically, this concept will be demonstrated with x-ray computed tomography images of voids constituting a database of hundreds of thousands of possible microstructures. At each material point in a component, the specific microstructure is chosen based on the results of a process model for AM. Finally, the load history during the component's expected service life is predicted, and used to estimate the fatigue life (or another performance indicator).

By using a database of microstructures and conducting a sampling study (viz Monte Carlo analysis), the need to define a representative volume element is alleviated, albeit at a computational expense. We can also enable prediction of performance variability and distribution by introducing some randomness into the selection of each microstructure. The choice of microstrcture at each material point in the component is based on processing history, making it possible capture the difference between, e.g., different toolpaths or choice of processing parameters. In the effect, this appears something like a digital twin, where a given build might be tracked through to a failure profile. One might alternatively think of this a “virtual experiment,” as each “specimen” tested computationally provides one data point in terms of performance (e.g fatigue life), much like a physical experiment.

The resulting framework has several unique features: location- and history-dependent properties and performance prediction, scalability to components or even systems (based on the macro-scale solution method employed), and ability to predict variability, in addition to mean/min/max, in behavior throughout the domain

This comes at a cost, naturally. A sufficiently rich database of images of microstructures ought to be used; “sufficiently rich” is ambiguous and depends on many factors. Some way to connect processing history to a suitably measure of microstructure that is both location/processing dependent, predictable, reliable, and relevant to the mechanical performance is also optimal.

Metal AM for use in structural applications where fatigue loading might occur is an excellent challenge with which to demonstrate this method. In AM, relatively unique, dispersed, heterogeneous microstructural features arise that depend upon a host of factors related to the conditions under which a part is fabricated. A truly predictive model can provide confidence in the robustness of designs and a quantifiable safety assessment, while minimizing the number of experiments required. However, this is only possible when using a model with sufficient descriptive capability.

Understanding the mechanical properties of AM metals has advanced rapidly in the past few years. Many of these advances have been experimental, and predictions of the properties of AM materials such as Ti-6Al-4V, SS316L, and Ni-based superalloys have been reported. One commonality between these materials is that they may widely vary point-to-point within a single component, between builds with different parameters, and between different builds on the same machine with the same conditions and parameters; the challenges associated with these various sources of variability have been noticed by previous computationalists. Experimental efforts have noted this variability, too; for example, Gong et al. and Sheridan et al. show significant variance between builds with different parameters and within builds with the same parameters, highlighting the importance of processing parameters, but also process-induced randomness. This kind of variability and randomness, particularly where defects are concerned, results in heterogeneous material for which standard prediction methods may be ill-suited.

Multiscale modeling is one approach that might be able to capture heterogeneous, non-uniform material responses. For example, Horstemeyer provides a case of fatigue modeling in heterogeneous materials using a hierarchical multiscale approach based on the multi-stage fatigue model developed. More recent works have focused on applying models throughout the processing and subsequent service life for additive manufacturing to relate processing, structure, and performance. These frameworks are valuable, and although not multiscale, and provide a starting place for the current work. Both are relatively deterministic, and critically both use process modeling to directly predict microstructure and defects. This implies complete reliance upon the accuracy and veracity of the process model to capture all important physics—for defects such as pores an accurate prediction remains an challenge within the AM modeling literature, particularly at the component scale. When fatigue response depends on the precise shape and location of each pore, reliance on a model might be unwise.

In the following, a sketch of the methods employed to build our multiscale model is made, including a thermal model of the AM process, an image- and modeling-based statistical description of voids, and the mechanical multiscale model used to predict performance. This is followed by a computational example, where prediction of the strain-life behavior of a sample of specimens built of Inconel 718 with several different process parameters is demonstrated.

Methodology

Conceptually, this multi-physics, multiscale method is composed of three major parts. The first two are used to generate the required information that describes the heterogeneity within the materials, and last uses this information to make mechanical performance predictions. As such, the first two are what relates the method to AM metals; the third part is the primary contribution and is not necessarily restricted to any particular material system.

The first part involves thermal modeling of the build process, so that the influence of processing parameters is captured. The second part uses synchrotron x-ray computed tomography images of voids in a prior AM build to map, based on the processing information in the first part, possible microstrctures to the component. A Monte Carlo type approach is used to generate many possible instantiations to account for random fluctuations within the same processing conditions (measured as scatter in the processing-structure relationship). The third part is a concurrent multiscale stress analysis that uses the instantiations from the second part and computational crystal plasticity to estimate the fatigue potency of voids on the micron-scale throughout a centimeter-scale component. A standard test specimen will be used as a consistent example throughout, although the method is widely applicable.

FIG. 89 shows an overall diagram of the computational scheme. Geometry, build process parameters, material, and loading conditions must be specified. These are used to conduct a thermal analysis and a macroscale stress analysis. For each material point X within these two models, an element-wise sub-model is constructed to represent a possible state at that point. This uses local thermal history and strain history to determine the microstructure (void geometry) and deformation history. These are used to predict the microscale evolution of state variables such as plasticity and damage, which are homogenized (e.g., by taking the I norm of the domain) and used as element-wise estimators of part-level susceptibility to failure.

Model Setup

The system described here is applicable to the same geometries, choices of processing conditions, boundary conditions during mechanical load, and materials that can be represented in standard finite elements analysis. Sufficient experimental data, especially 3D images of defects as will be described later, are required for the material of interest.

Macroscale

Thermal Modeling for the AM Build Process

We start by conducting a thermal analysis to model building the component of interest using the directed energy deposition (DED) method. This analysis is done using a transient thermal Finite Element solver. The governing heat transfer energy balance to be solved is:

ρ c p t = x i ( k T x i ) + Q ( 6 - 2 )

where ρ is the material density, cp is the specific heat, t is the time, xi are the spatial coordinates, k is the conductivity of the material, T is the temperature, and Q represents the heat source.

This heat source is represented by a moving laser described by the Gaussian distribution:

Q = 2 P η π R b 2 exp ( - 2 ( x 2 + y 2 + z 2 ) R b 2 ) ( 6 - 2 )

where P is the power of the laser, η is an absorptivity factor to limit the amount of energy absorbed by the material from the laser which was taken to be 30%, and Rb is the radius of the laser. The variables x, y, and z are local coordinates of the laser. Heat loss on the dynamic free surfaces of the model is simulated though a combination of convection and radiation. Convective heat loss is defined by


qconv=hc(T−T)  (6-3)

where hc is a convection coefficient, T is the surface temperatures, and T is the far-field (ambient) temperature. Radiation heat loss is defined using the Stefan-Boltzmann law, given by


qradsε(T4−T)  (6-4)

where σ is the Stefan-Boltzmann constant and ε is the surface emissivity of the material.

Particular build parameters, including laser speed and power and the toolpath are selected. The material must also be specified. With this information given, the model can predict the time-temperature-history of each point within the part. The solidification cooling rate (SCR) is calculated based upon the temperature history of the thermal model and outputted at each node. This is approximated, according to Eq. (6-5), as the time it takes for a material point, represented by subscript i, to reach the solidus temperature from the liquidus temperature. In order to capture this solidification behavior Eq. (6-1) is solved explicitly with an approximate timestep of 9.0×10−4 s. If too large of a timestep is chosen, one may skip over the solidification behavior at some material points. Additionally, in the case of re-melting, only the final cooling stage is considered.

S C R ( X M ) T X M l i q u i d - T X M solid t X M l i q u i d - t X M solid X M Ω M ( 6 - 5 )

where SCR(XM) is the solidification cooling rate as a function of the macroscale spatial coordinates XM within the macroscale domain ΩM, TXMsolid is the solidus temperature, TXMliquid is the liquidus temperature, tXMliquid is the time at which the liquidus temperature is reached, and tXMsolid is the time at which the solidus temperature is reached. However, more information is required before a mechanical model of the component can be conducted. Specifically, the anisotropic distribution of microstructure, including defects such as voids, developed during the build needs to be estimated.

Microscale

Relate Thermal Conditions and Defects

To use the thermal prediction, we need to connect thermal model outputs to defect statistics and geometry. We choose to generate this relationship in two stages: first through a relationship between a thermal descriptor (e.g., SCR, as computed above) and microstructure descriptor statistic, then through a database of microstructural geometries which correspond to each microstructural descriptor (e.g., voids size) statistic. The microstructures from this database is then used to populate each realization of the part. This defect estimation and database building process is outlined in FIG. 90. In the first part, a relationship between solidification cooling rate (SCR) and void volume fraction (Vf) is determined using process modeling and X-ray computed tomography. Panel (a) of FIG. 90 is the subsets of the images acquired with X-ray tomography are selected on the basis of Vf, such that the expected range of Vf for any arbitrary part (with known or predicted thermal history) is spanned. Panel (b) of FIG. 90 is a database of these possible microstructures is generated, including computing the training stage of the mechanical model.

In order to identify the SCR-to-void relationship and build a database, we used two Inconel DED single-track thin wall parts built with processing parameters corresponding to parameter set 1 in Table 6-2. Both walls used a vertical zig-zag toolpath pattern, but while one wall's toolpath was continuous, the other added a one-minute dwell between each layer.

X-ray tomography imaging experiments were performed at Beamline 2-BM at the Advanced Photon Source, Argonne National Laboratory on specimens extracted from 22 locations on the two thin walls (total 44 data points). Each image was of about 1 mm3 of material with voxel edge length 0.65 μm. Contrast from x-ray absorptivity was used to distinguish between voids and material using a series of processing steps including filtering, thresholding, and artifact removal. Eleven different descriptive statistics were extracted from these images, such as void location, size, shape, orientation, and n-nearest neighbor information. For simplicity, we will focus here on the void size, as represented by the single-point correlation statistic: void volume fraction Vf. The overall Vf was computed from the sum of the voids sizes throughout the image for each location.

The build process of these two thin walls is modelled using the thermal analysis method outlined in above. The parts' thermal history is summarized as a point-wise SCR throughout the build. The SCR is a useful, physically relevant single-point statistic that summarizes the thermal conditions that result from the choice of building parameters. The average SCR at the location of each of the x-ray images was computed using this processing model. Note that experimental measurements of the cooling rate, e.g., with an infrared (IR) camera, could be used to provide equivalent data. The first part of FIG. 90 shows how the average SCR versus the Vf of each image is gathered and plotted. An exponential relationship given by:


Vfrac=Ae−(B)(SCR)  (6-6)

is fit to the data. The fit parameters are A=0.0047 and B=0.0011.

Build Image Database

Next, this relationship is used to determine a range of possible microstructures that might occur within the build. Subsets of the images used in the first part were exhumed, such that the Vf range of the subsets corresponds to the range estimated to occur in a part as given by the relationship shown in FIG. 90. These image subsets constitute a database of possible microstructures, in this case each was of size 97.5 μm×97.5 μm×97.5 μm and between about Vf=0.0001 and Vf=0.03.

To complete the database, the three training steps outlined in Sect. 2.4.1 were conducted on each entry. Thus, the final database used for the multiscale mechanical response prediction contains proto-data used to generate response predictions for arbitrary loading, which depends upon microstructural information.

Mechanical Response Prediction

In the previous two sections, we have used a database of experimental data to populate the part with a realistic, heterogeneous distribution of voids that correspond to the processing conditions used to build the part. The next step is to use this information to predict the mechanical properties of the part. However, for the same reason that it was computationally infeasible to directly predict the defect structure, it would be infeasible to directly model the mechanical response of the microstructures we have generated.

In order to capture the microstructural information required to predict mechanical performance of AM materials, multiscale approach is taken. At the macroscale, a simple Johnson-Cook material model is used to provide the strain boundary conditions (more specifically, deformation gradients are used) to the data-driven reduced order model used at the microscale. At the microscale, crystal plasticity is used to predict the material behavior. Lacking more complete data, we simply assume that any microstructure with Vf within ±10% of that specified by Eq. (6-6) has equal probability of occurring. In this way, the part is described by essentially a Monte Carlo process, where the possible states of the random variable is controlled by the process model. A schematic of this process is shown in FIG. 91, where the local temperature history and strain history are taken as inputs to a micromechanics model that reports a homogenized response. For each macroscale material point (element, in this case), the thermal history and strain history are passed to a microscale solver; a microstructure is selected from the database developed based on the thermal history, and deformation boundary conditions are applied according to the strain history. A crystal plasticity based microscale solution is computed, and a homogenized response (e.g., the l-norm of the fatigue indicating parameter, if a fatigue problem is chosen) is returned to the macroscale.

Microscale Reduced Order Model with Crystal Plasticity

At the microscale, the modeling approach combines data-driven micromechanics with computational crystal plasticity, termed crystal plasticity self-consistent clustering analysis (CPSCA). The method is derived from first order homogenization, the Hill-Mandel condition, and local equilibrium. This problem is defined concisely in a finite deformation as:

( P X = 0 , X Ω , F = u X , X Ω , 1 Ω Ω F d Ω = F 0 . ( 6 - 7 )

where P is the first Piola-Kirkchoff stress, u is the displacement at point X in domain Ω, F is the displacement gradient, and F0 is the remove (applied) deformation.

To satisfy the Hill-Mandel condition we assume a periodic displacement field within the microscale domain and anti-periodic boundary traction. Under this assumption, the boundary value problem given in Eq. (6-7) has been shown to be equivalent to the Lippmann-Schwinger equation, and approximated clusterwise as


FIJ=1NcDIJ:[PJ−C0: FJ]−F0=0, with I=1, . . . ,Nc,  (6-8)

where the domain has been discretized into a fixed, finite number of clusters Nc, C0 is a reference stiffness, and * denotes convolution, and the tensor DIJ describes the interaction between clusters I and J, given as

D I J = 1 | Ω I | Ω χ I ( Γ 0 * χ J ) d Ω . ( 6 - 9 )

where ΩI is the domain in cluster I, Γ0 is a periodic fourth order Green's operator, and χ is the characteristic function. For a detailed derivation.

CPSCA solves this equation in two stages. The first “training” stage includes three parts: data collection, data compression (or clustering), and computation of the interaction tensor. The resulting interaction tensor can be stored for future use. The second “prediction” stage makes use of the interaction tensor and solves the integral equation given in Eq. (6-8) subject to an applied strain state and material law.

This first stage is conducted for each of the subset volume, adding the clustering and interaction tensor data to the database of image subsets.

Thus, after a part is instantiated and subsets used to populate the microscale, only the second stage has to be used compute the mechanical response. The second stage can then be run any number of times, with independent stress-strain histories and boundary conditions. In the second stage, cyclic loading is applied, and the stresses are computed with a crystal plasticity (CP) material law. The applied deformation gradient is decomposed into an elastic and plastic part; the plastic part of the deformation gradient is computed from the plastic velocity gradient, which itself is determined by summing the plastic shear velocity across slip systems in the intermediate configuration, as:


{tilde over (L)}pα=1Nslip{dot over (γ)}(α)({tilde over (s)}(α)⊗ñ(α))  (6-10)

where ⊗ is the dyadic product, α is a slip system, Nslip is the number of slip systems, {dot over (γ)}(α) is the microscale shear rate, s(α) is the slip direction, and n(α) is the slip plane normal. The phenomenological power-law with backstress shown in Eq. (6-11) is used to update the shear slip rate.

γ . ( α ) = γ . 0 τ ( α ) - a ( α ) τ 0 ( α ) m sgn ( τ ( α ) - a ( α ) ) , ( 6 - 11 )

where τ(α) is the resolved shear stress (computed with τ(α)=σ:(s(α)⊗n(α)) on slip system α, {dot over (γ)}0 is a reference shear rate, τ0(α) is a reference shear stress, a(α) is a backstress term, and m is a “rate hardening” coefficient.

Example Performance Measure: Fatigue Life

Failure prediction encompasses a range of damage mechanisms, which depend on predictions of properties. One of the more challenging failure to attempt to predict is that caused by cyclic loading: fatigue. Thus, we demonstrate this method with an application to measuring fatigue performance. Specifically, to predict fatigue crack incubation life, a fatigue indicating parameter (FIP) derived from the critical plane approach is used to estimate the fatigue incubation life of the microstructure, given a plastic strain history. The FIP is defined by

F I P = Δ γ max p 2 ( 1 + κ σ n max σ y ) ( 6 - 12 )

where Δγmaxp is the maximum cyclic plastic shear strain, σnmax is the peak stress normal to the plane on which ΔγmaxP occurs, σy is the yield stress, and K is a normal stress factor assumed to be 0.55. The FIP is related to number of incubation cycles using:


NFIPmax=γf(2Ninc)c  (6-13)

where γf and c are multiplicative and exponential Coffin-Manson-like calibration factors.

Numerical Demonstration

Calibration/Validation Case with Standard Fatigue Specimen

To demonstrate the method, a fatigue specimen conforming to the ASTM E606/E606M/E466 standard geometry is numerically tested. The specimen geometry was meshed with two different hexahedral meshes, for one thermal and one stress analysis, as shown in FIG. 92. The two meshes are largely the same, except the mesh used for stress analysis is coarsened in the thickness (z-) direction. This is the build direction, and thus requires at least one element per build layer in the thermal model, but during the stress analysis the stress and strain are roughly constant through this direction. This may, however, smooth out some of the variability between process parameters, as SCR was averaged during mesh coarsening. Following the schematic shown in FIG. 89, this mesh was loaded into both the thermal solver then the stress solver. An in-house thermal FEA code was used for the thermal model, C3D8R elements in ABAQUS were used to compute stress and strain at the macroscale. The lower grip was fixed and the upper grip was displaced at constant strain rate over five fully reversed load cycles at several different strain amplitudes.

Macroscale homogeneous material properties corresponding to Inconel 718 (IN718) were applied to both the thermal and macroscale stress analyzes. These thermal properties for IN718 are summarized in Table 6-1. The Johnson-Cook parameters were used, although this choice is essentially arbitrary—under the fatigue loading specified, the macroscale response was entirely elastic.

TABLE 6-1 Thermo-physical properties of IN718 Property Notation Value Density ρ (kg/m3) 8100 Solidus temperature Ts (K) 1533 Liquidus temperature Tl (K) 1609 Specific heat capacity cp (J/kgK) 360.24 + 0.026 T − 4 × 10−6 T2 Thermal conductivity k (W/mK) 0.56 + 2.9 × 10−2 T − 7 × 10−6 T2 Latent heat of fusion L (kJ/kgK)  272

For the thermal analysis, a zig-zag tool path with 90 layer-by-layer offset was selected, with the four different process parameters specified in Table 6-2. The part was meshed and simulated with the gauge section aligned normal to the build direction. A relatively fine mesh with 539,216 hexahedral elements (panel (b) of FIG. 92, substrate not shown) was used; panel (b) of FIG. 92 shows a detail of the mesh in the narrowest part of the gauge section. Panel (c) of FIG. 92 shows details of the specimen meshes, including the difference between stress (left) and thermal (right) meshes.

TABLE 6-2 Set of process parameters for thermal analysis of IN718 Parameter set Laser power Scanning Beam radius Layer (W) speed (mm/s) (mm) thickness (mm) 1800 15 1.5 0.75 1800 10 1.5 0.75 1500 15 1.5 0.75 1500 10 1.5 0.75

The CP model was used at the microscale. The parameters were calibrated by minimizing the difference between many runs of a cubic domain including 64 cubic grains and a set of baseline tensile and cyclic loading data for AM IN718, with starting conditions taken for m and {dot over (γ)}1 and for elastic moduli. The resulting the model parameters are given in Table 6-3. The parameters that relate FIP to fatigue life are fit to experimental high cycle fatigue data of IN718 collected from literature. We acknowledge that this generally represents relatively defect-free material, and may not be perfect for AM material; however, fatigue data for AM IN718 is relatively scant, and producing such data was outside the scope of this work.

The imaged voids identified by the method above were assumed to be embedded within a single crystal oriented so that the fastest growth direction was aligned with the build direction. This is consistent with experimental experience which suggests that grains in IN718 are much larger than the voids and preferentially orientated. Future work could use method, such as the cellular automata approach to predict grains throughout a build. Another possibility would be to synthetically generate grain structures from given experimental evidence which provides a statistical basis for the synthetic generation. Preliminary, as-yet-unpublished results indicate that grains generated this way could substantially impact predicted fatigue lives when used in concert with image-based void geometries.

A snapshot of the thermal response for the part being built under one of the processing parameter sets is shown as contours of temperature in FIG. 93, as an example of the thermal prediction. This thermal prediction progresses through the full building process for this specimen.

TABLE 6-3 Primary crystal plasticity model parameters Parameter Value C11, MPa 257,000 C12, MPa 127,000 C23, MPa 94,000 {dot over (γ)}0, s−1 0.0024 m 60 τ0, MPa 360

We simulated thirty different virtual test specimens, each with a different, random distribution of defects, and thus estimated fatigue life. This is shown in FIG. 6, where five realizations of specimens were simulated at three different load amplitudes for two different processing conditions. Each specimen includes 19,360 elements at the macroscale, each of which is represented by a 150×150×150 voxel mesh at the microscale. Thus, each specimen is represented, effectively, by about 65 billion voxels. Put another way, this 30 specimen test suite required an evaluation of 580,800 of these microscale volumes, or a total of 1.9602×1012 calls of the crystal plasticity routine over 130 time steps for each voxel. Without the CPSCA method used at the microscale, this would be a vast computational expense. However, with CPSCA, each virtual specimen test took about 9 hours using 36 cores in parallel on an Intel Xeon Skylake 6140 at 2.3 GHz clock speed.

FIG. 94 shows estimated fatigue lives for multiple realizations of the fatigue coupon, run at different applied strain amplitudes, mimicking experimental conditions. Two different processing conditions (conditions 1 and 2 in Table 6-2) were modeled. The results in FIG. 94 directly demonstrate the key features of this method. The contour plots of the gauge sections report an estimated number of cycles required to cause fatigue crack initiation at each macroscale point in the fatigue specimen; at any given macroscale point, the microstructure is different between different instantiations (Specimen A versus Specimen B in FIG. 94), which results in different contours plots. Comparing the microstructures between Specimen A and Specimen B provides the reason behind this difference: features with higher fatigue potency might, by random chance, occur at the point of highest strain concentration in one instantiation but not another. This is similar to behavior seen in physical testing. Although we assume that voids always occur, this is not a necessary assumption of the method—given sufficient characterization data of the processed material, many classes of microstructure might be seeded using this method.

This microstructures used here do not explicitly capture surface effects for the microstructural fatigue behavior. For example, a void on the microscale near surface of the entire part might have an impact of fatigue performance. While the macroscale effects of the boundaries are naturally included, this small scale interaction is a matter of ongoing work. Some authors suggest that there is limited change in overall fatigue life, up to the high-cycle limit (runout), for as-built versus machined finish specimens; this may indicate the assumption made here is reasonable.

A simplifying assumption for the grain structure was also used for this demonstration; however, we could add a step that either predicts the grain structure from thermal history, as was demonstrated, or derives a grain structure from experiments should that data be available. Currently, the kind of 3D grains structures needed to include image-based grains are not available to us, although one might consider using statistically-similar, synthetically generated grain structures based on currently available images the next logical step in purely image-based microstructures.

An implicit assumption of the images used here is that the track spacing was appropriate to avoid lack-of-fusion defects between tracks. This is because the images used to make the database came from single-track, thin-wall build, were no between-track porosity would be possible. This assumption could easily be relaxed by including images of voids in multi-track builds in the database.

Development of a benchmark for fatigue prediction in AM would support these modeling efforts. Currently, conflicting reports of the influence of AM, versus conventional processing, on the fatigue properties of metals exist. This can in part be attributed to the wide range of materials, but also to a range of build processes and choices made by machine operators. Without a better, AM specific, standard test procedure such conflicts will likely persist, to the detriment of the modeling community who lack calibration and validation data.

This example presents a method that exploits computational efficient micromechanics techniques to perform Monte Carlo-style numerical experiments. The specific microscale solutions are derived from a cluster-based solution of the Lippmann-Schwinger equation, and involve the prediction of fatigue life using crystal plasticity and a fatigue indicating parameter. A database of possible microscale geometries is developed from 3D imaging experiments. These geometries are related to AM processing conditions through the solidification cooling rate, and a process- and microstructure dependent, stochastic prediction of fatigue life is achieved with reasonable computational expense.

Example 7 Efficient Multiscale Modeling for Woven Composites Based on Self-Consistent Clustering Analysis

In this exemplary embodiment, the curse of computational cost in woven RVE problem is countered using the SCA, which maintains a considerable accuracy compared with the standard FEM. The Hill anisotropic yield surface is predicted efficiently using the woven SCA, which can accelerate the microstructure optimization and design of woven composites. Moreover, a two-scale FEM×SCA modeling framework is proposed for woven composites structure. Based on this framework, the complex behavior of the composite structures in macroscale can be predicted using microscale properties. Additionally, macroscale and mesoscale physical fields are captured simultaneously, which are hard, if not impossible, to observe using experimental methods. This will expedite the deformation mechanism investigation of composites. A numerical study is carried out for T-shaped hooking structure under cycle loading to illustrate these advantages.

Woven composites are widely used in industries such as aerospace and automotive because of their excellent mechanical performances as compared to unidirectional laminated composites. However, performing structural analysis of woven composites is challenging due to the mesoscale and microscale heterogeneities (see FIG. 95). Different features can be observed at these different scales, and simply homogenizing the composite structure and applying phenomenological constitutive relations that only characterize the average behavior of the material does not account the localized behavior at the finer scales. As a result, local nonlinear deformation and damage effects are not considered. In addition, the macroscale properties cannot be predicted based on the microstructural constituents, and experiments are required to design new composites, which are costly and time consuming.

Multiscale simulation provides a powerful method for analyzing both the material microstructure and macrostructure. Using this method allows the macroscale performance of woven composites can be predicted based on the properties of the constituents. Once the microstructure is characterized, macrostructural experiments are not needed every time the microstructure is changed. This allows the multiscale method to accelerate material design of woven composites, reduce the cost, and improve the analysis accuracy of woven composite structures. Moreover, the physical fields in different scales can also be captured, which are hard, or sometimes impossible, to observe using experimental method. Accomplishing effective multiscale simulations for woven composites still involves some challenges.

The first challenge is to find an efficient woven RVE solution. Effective macroscale properties are homogenized properties of composites, which are always adopted for the material selection and structural design with woven composites. To predict these effective properties, an RVE for the woven composite material must be developed, which will establish the link between the microstructural features and effective macrostructural properties. In the case of a periodic woven architecture, a unit cell is used for the RVE. For the microstructure design, the woven RVE solution can be integrated into an optimization algorithm in which the RVE has to be solved repeatedly to find the optimized solution and satisfy the requirement of objective effective properties. Therefore, solving the woven RVE problem efficiently can accelerate the whole process of optimization. As a result, it will promote the microstructure design of woven composites. Currently, several approaches have been proposed for solving the RVE problem. The analytical approaches, such as mixtures rules and theoretical micromechanics methods are efficient, but will lose accuracy in the case of complex microstructure and nonlinear, history-dependent material laws. The Direct Numerical Simulation (DNS) method, such as FEM, is extremely time consuming. The FFT-based method is more efficient than FEM, but encounters convergence problems for the high phase contrast in nonlinear problems. The Transformation Field Analysis (TFA), the Nonuniform Transformation Field Analysis (NTFA) and Proper Orthogonal Decomposition (POD) are other solution methods, but they require extensive a priori simulations to obtain deformation modes, especially for nonlinear phase behavior.

The second challenge is concurrent multiscale simulation for woven structures. The behavior of woven composite structures is predicted using the behavior of the RVEs through the concurrent multiscale simulation. Additionally, the physical fields in different involved scales can be captured simultaneously, which will expedite the deformation mechanism investigation of woven composite structures. Concurrent simulation requires numerous RVE solutions, which is computationally expensive using the FE2 framework, as shown in FIG. 96. In this example, only 5000 elements are used at the macroscale level, 1,843,200 elements are used at the woven RVE mesoscale level, and 576,000 elements are used at the UD RVE microstructural level. For the concurrent multiscale simulation, every up-scale material point will be linked with a down-scale RVE. In this example, assuming a single integration point for each element, 2,511.4 trillion elements are required for the entire multiscale computation. This computation is extremely expensive, and would require the use of a High-Performance Computing Cluster (HPCC).

Solving these challenges require improving the solution efficiency of the RVE problem. The SCA is an effective and efficient method to solve the RVE problem, which can be used for complex woven architecture undergoing irreversible processes, such as inelastic deformation. This makes it particularly attractive for integration into a multiscale simulation. The SCA method involves a two-stage process, an offline stage and an online stage. In the offline stage, a clustering algorithm is used to reduce the overall degrees of freedom (DOF) of the RVE, resulting in a reduced order RVE. In the online stage, the reduced order RVE is utilized for solving the discrete incremental Lippmann-Schwinger integral equation to obtain the stress and strain fields in the reduced order RVE. This efficient method has been used for simulation for 2-dimensional (2D), two-phase composites, and 3-dimensional (3D), hard inclusion material considering nonlinear, elastoplastic damage softening effect and computation for polycrystal material. These simulations have demonstrated good efficiency and accuracy.

In this exemplary example, the reduced order modeling process of woven composites by SCA is discussed and the results are compared with FEM. Moreover, the multiscale framework of woven composites is presented for a woven composite. Based on this framework, the part scale mechanical response, whether linear or nonlinear, can be predicted efficiently only using the fiber material and matrix material laws.

Methodology and Framework

SCA Method for a Woven RVE at the Mesoscale Level

A woven composite material is constructed by interweaving yarns in two directions and then filling the weave with an epoxy matrix material. The effective elastic properties of an individual yarn are predicted using a UD RVE based on the constituent properties of the fiber and matrix materials (see FIG. 97). Then, the woven RVE is meshed by high-fidelity voxel elements and the elastic analysis is conducted to obtain strain concentration tensor in each element. The degrees of freedom in the woven RVE domain are reduced by clustering these voxel elements based on the strain concentration and orientation in each element. Using the results of the woven RVE clustering, a material database is generated using the method that includes the interaction tensors DIJ, the strain concentration tensor of each cluster AI, the volume fraction cI, and the material parameters of the constituents. In the online stage, a Newton-Raphson iteration algorithm is adopted to solve the discrete incremental Lippmann-Schwinger integral equation set, which can improve the accuracy and convergence, especially for nonlinear material behavior. The solution is the mesoscale strain and stress fields.

FEM×SCA Concurrent Multiscale Framework

Two scales, the macroscale and mesoscale, are utilized in the concurrent multiscale framework in the study (FIG. 98). The structural scale (macroscale) is discretized by FEM, which can adapt to complex geometries. The woven RVE scale (mesoscale) is modeled using SCA. A multiscale simulation involving both the macroscale and mesoscale levels is performed in which the information is exchanged concurrently.

The load is applied to the structural scale model. At each integration point in the macroscale elements, the strain increment will be passed from the FEM model in the SCA model. This strain increment is applied to the corresponding woven RVE, and the SCA method is used to solve the woven RVE problem and return the stress increment to the FEM solver. The algorithm for the the concurrent multiscale simulation of woven composite structures is summarized as follows.

    • 1. Mesh the macroscale woven composites part using FEM;
    • 2. Begin solution increments;
    • 3. Compute integration point field variable from nodal values;
    • 4. for i=1, N_IP (Loop over integration points);
      • a. The macroscale strain increment ΔE is passed to user-defined subroutine as input data;
      • b. Run online part of SCA to solve woven RVE subjected to ΔE using offline woven database;
      • c. Compute the mesoscale strain increment Δe in every cluster in woven RVE domain;
      • for j=1, N_CLU (Loop over all clusters in this corresponding woven RVE);
        • Compute mesoscale stress increment Δσ using corresponding material model;
      • end for (Obtain the response at all clusters);
      • d. Check convergence of the reduced-order discrete incremental Lippmann-Schwinger integral equation, if not, update Δe using Newton-Raphson method and go to c, if yes, go to e;
      • e. Compute macroscale stress increment ΔΣ by averaging Δσ in woven RVE domain and pass the macroscale stress back to the FEM solver;
    • 5. end for (Obtain the response at all integration points); and
    • 6. Check convergence of the FEM part, if not, update nodal values and go to step 3.

From the flowchart, the SCA online algorithm can be implemented by the user-defined subroutine, which can be integrated with most commercial FEM software packages. In this way, the FEM×SCA multiscale framework can be adapted to arbitrary structural geometry and arbitrary woven architecture. Note that the cluster geometry is not required to be regular, which makes it effective for complex microstructure.

Verification of SCA for Woven RVE

Geometry Model

In the family of woven composites, plain weave composites are widely used for ease of manufacturing. A plain weave composite is selected as an example to demonstrate and verify the SCA method at the RVE level. FIG. 99 shows the plain weave RVE microstructure used in the present work. The cross section of the yarns is assumed to be elliptical, and the centerline of the yarn is modeled as a sine function. Two coordinate systems, X1O1Y1 and X2O2Y2, are created to describe the cross section and yarn centerline features, respectively. The mathematic description can be shown in Eq. (7-1). The local coordinate frame 1-2 is used to indicate the local orientation of the yarn, which is also the local system the material model in yarn. The 1-direction is the tangent of the centerline, and the 2-direction is normal to the 1-direction. This local frame varies along the length of the yarn. The woven RVE used in this example has 120 voxel elements in both width and length dimensions, and 32 voxel elements in height dimension.

The cross sectional shape and longitudinal shape are respectively modeled as

{ x 1 2 a 2 + y 1 2 b 2 = 1 y 2 = A sin ( 2 π l 0 x 2 ) ( 7 - 1 )

where a is long axis, b is short axis of the elliptical cross section respectively, A is amplitude of the sine function in Eq. (7-1). The three-dimensional geometric model of the woven RVE is defined by five parameters, which are also shown in FIG. 99. The values of these parameters are assumed and only for numerical verification purposes. With proper experimental characterization, it is possible to generate a realistic woven RVE with these parameters.

Material Properties and Constitutive Model

A nonlinear epoxy plastic material model is used to model the polymer matrix. The yield function is written as


ƒ(σ,σct)=6J2+2I1c−σt)−2σcσt  (7-2)

where σ is Cauchy stress tensor, I1=tr(σ) is the first invariant of Cauchy stress tensor,

J 2 = 1 2 η : η

is the second invariant of deviatoric stress

η = σ - 1 3 I 1 ,

σt and σc are yield strengths in tension and compression. A non-associative flow rule is used, with the plastic potential function written as


g(σ,σct)=6J2+2αI1c−σt)−2σcσt  (7-3)

where

α = 1 - 2 v plas 1 + v p l a s ,

νplas is known as plastic Poisson's ratio. Thus, the flow rule is given by

? = ? g σ ? indicates text missing or illegible when filed ( 7 - 4 )

where γ& represents the time derivative of the plastic multiplier. The evolution of yield strengths in tension and compression are written as


σtt0+Ht(1−e−ntα0)


σcc0+Hc(1−e−ncα1)  (7-5)

where σt0 and αc0 are the initial yield strengths in tension and compression, Ht and Hc are hardening parameters in case of tension and compression respectively. These material parameters are given in Table 7-1. α0 and α1 are internal kinematic variables, which are determined by the epoxy experimental data.

TABLE 7-1 Material parameters of matrix Parameter Value E(GPa) 3.76 v 0.39 vplas 0.3 σt0 (MPa) 29 σc0 (MPa) 67 Ht (MPa) 67 Hc (MPa) 58 nt 170 nc 150

A transversely isotropic elastic material model is considered for the fibers. In addition, the elastic properties are list in Table 7-2. The fiber volume fraction is assumed to be 60%. The subscripts 1, 2 and 3 indicate the local material orientation (see FIG. 99). The UD RVE (FIG. 100) used in this example has 240 voxel elements in both width and length, 10 voxel elements in height.

TABLE 7-2 Material parameters of fiber E1 (GPa) E2 = E3 (GPa) G13 (GPa) v12 v23 Fiber 231 12.97 11.28 0.3 0.45

The effective material properties of the yarn are predicted using a UD RVE (FIG. 100) model and the elastic properties (Table 7-. 1 and Table 7-2) by applying six orthogonal loads with periodic boundary conditions (PBC). As a result, the elastic material properties of yarn are presented in Table 7-3.

TABLE 7-3 Predicted effective elastic material properties of yarn E1 (GPa) E2 = E3 (GPa) G13 (GPa) v12 v23 Yarn 138.8 7.08 4.49 0.25 0 31

Clustering Process for the Woven Mesoscale RVE

The matrix material is isotropic, which requires that the clustering only be conducted once, based on the Am tensor. After this procedure, the material points with the most similar Am tensor will be grouped into the same clusters. FIG. 101 shows the clustering results of the matrix for 256 clusters using k-means clustering.

The clustering process for the yarn material will be more complex, as illustrated in FIG. 102, where clustering process and results of yarns with 64 clusters are shown. For each yarn, clustering is performed first based on local orientation. The resulting clusters are refined further using strain concentration Am tensor. Since one cluster corresponds one orientation dependent material law and the local frame aligns with the points in the yarn centerline (see FIG. 99). The 1-direction is tangent to the yarn centerline and represents the yarn material orientation for each material point. The clustering process is conducted twice. First, a single yarn is clustered based on the material orientation using k-means. The points with the closest material orientation will be grouped into the same cluster. Based on the results in the first step, the material points in the same cluster will be clustered a second time according to strain concentration tensor Am. In this exemplary example, two clusters are used the second time. After the two-step clustering, the material points with the closest orientation and the closest strain concentration tensor will be grouped into the same cluster. These two steps are repeated for all yarns in the RVE.

Results and Discussion

Uniaxial tension and pure shear responses are calculated using the SCA method. FIG. 103 and FIG. 104 include the stress-strain curves for the woven composites under these two different load cases. The number of clusters in the matrix ranges from 64 to 256, while the number of clusters in the yarns is fixed at 128. When the number of clusters in the yarns changes from 32 to 128, the number of clusters in the matrix is fixed to 256. The results from the FEM are also provided as comparison to the baseline solution. It can be concluded that the SCA results are in very good agreement with the FEM results under the pure shear condition in the linear and nonlinear regions. For the uniaxial tension condition, the SCA results converged with the FEM results as the number of clusters in the matrix increased. Since the nonlinear constitutive model is only used for the matrix material, more clusters are needed to capture the local material nonlinear effects. Compared with the computational cost of FEM, SCA is capable of accurately capturing nonlinear behavior of woven composites with significantly fewer degrees of freedom (SCA results differ by less than 4% compared with FEM). Table 7-4 and Table 7-5 present the efficiency comparison between the FEM and SCA methods. FEM required about 1.45 million DOF be solved with a solution time of 6523 seconds, while SCA only requires 2304 DOF and has a solution time of 30 seconds. As a result, the SCA method significantly improves computational efficiency in terms of time and memory requirements. FIG. 103 shows the prediction results given by FEM and SCA (The SCA-64-128 indicates 64 clusters in matrix and 128 clusters in the yarns).

TABLE 7-4 Computation efficiency comparison of FEM and SCA Elements/ Computational Time/s Clusters DOF Uniaxial Tension Pure Shear FEM 460800 1.45 million 6785.8 6523.1 SCA-64-128 192 1152 11.5 10.4 SCA-128-128 256 1536 18.2 17.8 SCA-256-128 384 2304 28.8 19.8 Max. Speed Up 2400 1258 590.1 627.2

TABLE 7-5 Computation efficiency comparison of FEM and SCA Elements Computational Time/s Clusters DOF Uniaxial Tension Pure Shear FEM 460800 1.45 million 6785.8 6523.1 SCA-256-32 288 1728 20.9 14.5 SCA-256-64 320 1920 22.6 17.1 SCA-256-128 384 2304 28.8 19.8 Max. Speed Up 1600  839 324.7 449.9

Property Prediction and Concurrent Multiscale Simulation

Macroscale Anisotropic Yield Surface Prediction

A nonlinear epoxy elastic-plastic material law is considered for the matrix, which results in the overall elastic-plastic behavior for woven RVE. The yield stress of the elastic-plastic material is an important property for material selection and design of composite structures. A yield surface is developed to evaluate material yielding under various loading conditions. The anisotropic Hill yield criterion is considered in this example for woven composites. The homogenized material law can be efficiently predicted using SCA based on the epoxy elastic-plastic material law for the matrix and the elastic material law for the yarn. The quadratic Hill yield criterion has the following form:


Fyy−σzz)2+Gzz−σxx)2+Hxx−σyy)2+2yz2+2zx2+2xy2=1  (7-6)

where F, G, H, L, M, N are constants characteristic of the yield surface, which are traditionally determined by burdensome experiments. Additionally, some experiments are difficult to perform, such as the out-of-plane tension test. In this exemplary example, these parameters are predicted using the SCA method, which significantly reduces the computational cost and improves the efficiency. If Yxx, Yyy, Yzz are the tensile yield stresses in the principal anisotropic direction, it can be shown that:

1 Y x x 2 = G + H , 2 F = 1 Y y y 2 + 1 Y z z 2 - 1 Y x x 2 1 Y y y 2 = H + F , 2 G = 1 Y z z 2 + 1 Y x x 2 - 1 Y y y 2 1 Y z z 2 = F + G , 2 H = 1 Y x x 2 + 1 Y y y 2 - 1 Y zz 2 ( 7 - 7 )

If Yyz, Yzx, Yxy are the yield stresses in shear with respect to the principal axes of anisotropy, then

2 L = 1 Y y z 2 , 2 M = 1 Y z x 2 , 2 N = 1 Y x y 2 ( 7 - 8 )

By taking advantage of the symmetrical features of the woven RVE, only four orthogonal loading conditions are applied to the RVE; the responses are calculated using the SCA method. The tangent stiffness is computed at each point from the stress-strain response, and the yield points are identified by evaluating the change in the tangent stiffness. As a result, the values of yield stress in six directions are obtained. In addition, the Hill constants in Eq. (7-6) are calculated using Eq. (7-7) and Eq. (7-8) using the values of yield stress. The six-dimensional yield surface descripted by Eq. (7-6) can be difficult to visualize, but by selecting three components at a time (and setting other three components to be zero), this six-dimensional yield surface is plotted in three-dimensional space, see FIG. 105 showing Hill yield surface calculation workflow. The 3D yield surfaces are plotted against three normal stress components and three shear stress components. For the plot against normal stress components, the cross section where sigma_zz=0 is illustrated. For the plot against shear stress components, the cross section where sigma_xy=0 is plotted.

For the nonlinear computation of woven composite structures, this anisotropic Hill yield surface can be used as a criterion to instantaneously identify the onset of the plastic deformation under various loading conditions.

The present workflow shown in FIG. 105 allows one to build yield surface for various microstructure and material constitutive information with minimum efforts (around one minute using a personal computer). A large woven composite response database can be built to assist design of woven composite against yielding. Given a priori information on maximum service loads, the database will provide all possible woven microstructure (e.g., yarn geometry and yarn angle) and material constituents (e.g., matrix properties and yarn properties) that would prevent yielding to occur. Hence, the workflow could potentially accelerate the woven composite design process by narrowing down the design space of various design parameters.

Multiscale Simulation Convergence Study

An RVE convergence study is first conducted to quantify the effect of RVE size on the stress-strain response (FIG. 106). RVE-1 is a unit cell of plain weave woven composite, and RVE-2 is eight times bigger than the RVE-1. The results for these two different RVE sizes are shown in FIG. 106. It is noted that the results are in close agreement with each other, and that the RVE-1 will provide converged results with greater computational efficiency.

T-Shaped Hooking Structure Analysis

Woven composites are generally made of multiple layers for industrial application. A T-shaped hooking structure is a common geometry for connecting different composite parts. In this example, multiscale simulation is used to capture the macroscale and mesoscale fields in different layers during cyclic bending of the T-shaped hooking structure. The structure and the loading condition are depicted in FIG. 107A. Multiple layers are considered through the thickness. The red highlighted area in FIG. 107B represents a critical zone when failure stresses are reached, as demonstrated through experiments, and a finer mesh is used in this area. The total number of elements is 34,720 for the structural level model.

For woven mesoscale RVE, 64 clusters in the matrix and 32 clusters in the yarns are considered for the SCA calculation, while the eight-node continuum brick element with a reduced integration (ABAQUS element C3D8R) element is used for the FEM calculation. The macroscale behavior is determined by the microstructural morphologies and the mesoscale constitutive equation of each cluster. The SCA material database is first generated during the offline stage, which makes the multiscale simulation more efficient.

This numerical study is implemented with an ABAQUS VUMAT User Subroutine and the discrete incremental Lippman-Schwinger equations are solved using Intel Math Kernel Library (MKL) FORTRAN codes. This numerical example is run on Intel® Xeon® processor with 48 cores and 128 GB memory

The computational results are presented in FIG. 108. Four elements in different layers around the corner are selected to present the mesoscale fields. For the bending loading condition, a stress gradient exists through the thickness, and the stress fields are different in different RVEs. Since the yarns have a much higher modulus, they undertake much more loading than matrix. The homogenized stress-strain curve at maximum stress location is plotted in FIG. 108, which shows the residual plastic strain after loading and unloading. The stress state at peak point is plotted on the 2D yield surface, which shows that the stress has already exceeded the initial yield surface. Additionally, the computational times are presented in Table 7-6, which demonstrates the significant improvements in efficiency of FEM×SCA framework.

TABLE 7-6 Computational time comparison Concurrent multiscale framework Computational time FE2 5.2 × 105 days (Estimated) FEM × SCA 2.4 days Speed up 2.16 × 105

The above numerical study presents the advantages using the proposed multiscale simulation framework. The stress and strain fields can be captured in both macroscale and mesoscale, including the nonlinear effects, which are difficult to observe using experimental technology. As a result, this framework establishes the connection between the microstructure and macroscale response of the composites structure. When the woven microstructure is modified, but the yarn and matrix material remain unchanged, no additional experiments are needed to calibrate the constitutive equations; only the SCA offline database needs to be updated. In this way, it reduces the cost and improves the efficiency to find the optimal microstructure for the specific structures. Given the efficiency of the SCA online, the larger dimensional composites structures can be analyzed using this framework.

In this exemplary example, a woven composite multiscale modeling framework based on Self-consistent Clustering Analysis (SCA) is established. A two-stage reduced order modeling for woven composite, represented as RVE, is developed: In the offline data compression stage utilized clustering technique to reduce the overall degrees of freedom in the RVE domain as material points with similar mechanical responses are grouped into clusters. An interaction tensor linking different clusters is computed afterwards, generating a woven RVE microstructure database; In the online stage, Newton-Raphson iteration solves the reduced-order discrete incremental Lippmann-Schwinger integral equation. It exhibits rapid convergence for both linear and nonlinear material laws.

The woven multiscale modeling approach provides two attractive features: 1) Given the woven microstructure, online stage can utilize different materials laws for matrix and yarn phases to compute woven microstructure responses. For example, for temperature dependent material properties, the woven behavior at different operation temperature can be computed efficiently. 2) Given the same constituents properties, one only needs to update the offline database to incorporate different wave structures, such as plain weave, twill weave, or satin weave.

The woven multiscale modeling framework has various potential application, where two important applications are illustrated in the present study:

Rapid yield surface generation for woven design against yielding. The yield surface generation workflow can be used to investigate whether the woven composite would be free of plastic deformation under possible loading conditions. Note that it can be easily extended to failure surface for woven design against failure.

FEM×SCA woven laminate modeling framework which captures macroscale (FEM mesh) and mesoscale mechanical behavior simultaneously during the analysis. The mesoscale field evolution can be tracked as the load increases and the bridge between microstructure and macro-response is built. Based on the FEM×SCA framework, the damage and failure model can be built in the mesoscale and take more mechanisms into consideration, even conducting composites structure level failure analysis. Compared to the traditional phenomenological constitutive relations, these mesoscale-mechanism-based constitutive relations do not need complex mathematical formulas and numerous parameters. In addition, this framework can be extended to larger scale structure level analysis with complex loading conditions. Finally, this work provides an efficient methodology and framework to solve the woven composites multiscale problems.

Example 8 Self-Consistent Clustering Analysis for Multiscale Modeling at Finite Strains

Accurate and efficient modeling of microstructural interaction and evolution for prediction of the macroscopic behavior of materials is important for material design and manufacturing process control. This study approaches this challenge with a reduced-order method SCA. It is reformulated for general elasto-viscoplastic materials under large deformation. The accuracy and efficiency for predicting overall mechanical response of polycrystalline materials is demonstrated with a comparison to traditional full-field solution methods such as finite element analysis and the fast Fourier transform. It is shown that the reduced-order method enables fast prediction of microstructure-property relationships with quantified variation. The utility of the method is demonstrated by conducting a concurrent multiscale simulation of a large-deformation manufacturing process with sub-grain spatial resolution while maintaining reasonable computational expense. This method could be used for microstructure-sensitive properties design as well as process parameters optimization.

The process-microstructure-property chain relationship plays an important role in the development of new materials, particularly in the practice of computational material design or integrated computational materials engineering (ICME), which relies heavily on microstructure-based models. These models, once calibrated, can be used to explore a larger material design space than traditional trial-and-error methods. Classical micromechanics include analytical models, for instance, the self-consistent method and the Mori-Tonaka method. These models are efficient but require stringent idealizing assumptions of microstructure morphologies and/or interactions. The last decades have seen a rise in detailed modeling of manufacturing process, microstructure evolution and resulting mechanical properties supported by the development of computational mechanics as well as powerful computers. However, the speed of ICME deployment is still limited by the computational complexity of the models involved. Therefore there is substantial interest in data-driven reduced-order models, which have the promise of providing high accuracy without the computational expense associated with the detailed models used heretofore.

These reduced-order models have two important features: (1) degrees of freedom (DoFs) are significantly reduced; and (2) some data are precalculated during offline stage so that it can be repetitively used for iterations during online stage. One example is the transformation field analysis (TFA) method. It decomposes local deformation into elastic deformation and transformation deformation (or inelastic deformation). The elastic deformation is determined by precalculated elastic strain concentration tensor. The transformation deformation is assumed to be uniform in each material phase. In this way, the DoFs are reduced to the order of the number of phases involved in a specific problem. Thus this method is faster than traditional full-field approaches, but fails to accurately capture intraphase heterogeneous mechanical response. Nonuniform transformation field analysis (NTFA) alleviates this problem by interpolating pre-calculated transformation field data (or modes) under some predefined loading paths; this achieves higher accuracy than TFA with DoFs on the order of the number of modes used. Other data-driven methods employ similar ideas to reduce DoFs, for example proper orthogonal decomposition (POD) interpolates the displacement field and calculates the modes more efficiently with single value decomposition. However, these methods share the same restrictions: many expensive offline calculations are needed to obtain representative modes for highly nonlinear material behavior such as plasticity and finite strain.

Recently, the SCA proposed by Liu et al. has been shown to maintain high accuracy and efficiency even with these more challenging loading conditions. To do this, SCA uses a clustering-based data compression technique for order reduction and a self-consistent iterative scheme to solve the Lippmann-Schwinger equation accurately. Some recent developments of this method include theoretical analysis of convergence of SCA, applications in toughness design of particle reinforced composites, damage process of elasto-plastic strain softening materials, and fatigue life prediction for a NiTi alloy. However, most of these studies are limited to small strain problems and only consider interacting microstructures with a relatively small number of features, such as homogeneous matrix, inclusions and/or voids. The finite strain SCA was used to model inclusion breakage during the drawing process. In practice, the constituents in the matrix phase should also be considered if their characteristic length is comparable to inclusions and voids. For instance, metallic alloys are composed of grains with various morphology, size, and lattice orientation as well as possible defects such as precipitates or voids. Although the self-consistent and eigen-deformation based methods have been used to model interacting polycrystals, their limitations, as outlined above, remain. We recently coupled a small-strain SCA formulation with a crystal plasticity model, but it neglects the rotation components of the deformation and thus is not suited to finite strains.

This exemplary study tackles the above limitations by generalizing SCA to the finite strain case and demonstrates its accuracy and efficiency for predicting macroscopic mechanical response of heterogeneous elasto-viscoplastic materials, e.g., polycrystalline materials. Enabled by this approach, two case studies that would involve impractically vast computational time otherwise are presented for a Titanium alloy: the first one quantifying a microstructure-property relationship with uncertainty, and the second one predicting texture evolution of a thick plate during rolling through concurrent multiscale simulation.

The following sections discuss the data-driven mechanistic SCA method for finite-strain MVE problems, the accuracy and efficiency of SCA under finite strains and compared to reference solutions using the finite element method (FEM) and the FFT method. The algorithm, implementation, and procedures used to generate our results are also given.

Finite-Strain Self-Consistent Clustering Analysis

Microstructural Volume Element Problem

The equilibrium mechanical response of a MVE under far-field macroscopic loading can be described by a set of equations formulated in the undeformed configuration:

( P X = 0 , X Ω , F = u X , X Ω , 1 Ω Ω F d Ω = F 0 . ( ( 8 - 1 )

In these equations, P is the first Piola-Kirchhoff stress (PK1 stress), F is the deformation gradient, u is the displacement, X is a material point, and Ω is the MVE domain. Pure deformation type far-field loading F0 is assumed here for simplicity.

Cluster-Based Lippmann-Schwinger Equation

Under the assumption of periodic boundary conditions, the MVE problem given by Eq. (8-1) is equivalent to the Lippmann-Schwinger equation


F(X)+Γ0*(P(X)−C0:F(X))−F0=0,∀X∈Ω  (8-2)

where Γ0 is the 4th order Green's operator associated with an arbitrary reference stiffness tensor C0 and * denotes convolution operation defined by


Γ0*(P−C0:F)=∫ΩΓ0(X−X′):(P(X′)−C0:F(X′))dΩ(X′).  (8-3)

To solve Eq. (8-2) numerically, a MVE domain decomposition is necessary. Unlike traditional numerical methods such as FEM and FFT, which do this by defining a fine (relative to the minimum feature size) mesh, SCA employs a clustering-based domain decomposition method to be introduced. Here it is assumed that the MVE is decomposed into Nc non-overlapping sub-domains, called clusters hereafter. For the Jth cluster ΩJ, J=1, . . . , Nc, the characteristic function is defined as

χ J ( X ) = ( 1 , if X Ω J 0 , otherwise . ( 8 - 4 )

Using characteristic functions, we approximate the deformation gradient and stress as


F(X)≈ΣJ=1NcχJ(X)FJ and P(X)≈ΣJ=1NcχJ(X)PJ,  (8-5)

where FJ is the average deformation gradient and PJ is the average stress in the Jth cluster. Thus Eq. (8-6) can be approximated as


ΣJ=1NcχJFJJ=1Nc0J):(PJ−C0:FJ)−F0=0.  (8-6)

Multiplying both sides of Eq. (6) with χI, I=1, . . . Nc, and integrating in Ω gives


ΩχIJ=1NcχJFJJ=1Nc0J):(PJ−C0:FJ)−F0)0.  (8-7)

The cluster-based Lippmann-Schwinger equation is obtained by simplifying the above equation:


FIJ=1NcDIJ[PJ−C0:FJ]−F0=0, with I=1, . . . ,Nc,  (8-8)

where DIJ is the interaction tensor between the Ith cluster and Jth cluster given by

D IJ = 1 Ω I Ω χ I ( Γ 0 * χ J ) d Ω . ( 8 - 9 )

Here |ΩI| is the volume of the Ith cluster.

SCA solves the cluster-based Lippmann-Schwinger equation in two stages. In the offline stage, the deformation concentration tensor field (known as the strain concentration tensor under the small strain approximation) is prepared and used to determine the clusters that define the regions ΩJ,J=1, . . . , Nc (domain decomposition), then the interaction tensors among these clusters are calculated. These data will be used in the online stage to solve Eq. (8-8) together with local material laws.

The application of the cluster-wise approximation made in Eq. (8-5) results in the loss of deformation compatibility. This means that although the solution of the continuous Lippmann-Schwinger equation, Eq. (8-2), is independent of the choice of reference stiffness tensor C0, the cluster-based approximation in Eq. (8-8) is not. To achieve higher accuracy, a self-consistent iterative scheme has been proposed to update C0 at each loading increment so that it approximates the effective tangent stiffness Ceff of the MVE. However, this necessitates updating the interaction tensor during the self-consistent iterative scheme, which would increase computation time in the online stage. Fortunately, most of the interaction tensor calculation effort can be done in the offline stage by enforcing C0 to be isotropi. The formulation of Ceff and its isotropic approximation are given below.

Offline Stage: Micromechanical Database, Clustering and Interaction Tensor Calculation

SCA reduces the degrees of freedom to be solved by taking advantage of the mechanical response similarity of material points in a MVE. This similarity is found by clustering the field data of some mechanical response. Generally, deformation concentration tensor can be used. It is defined by

A ( X ) = F ( X ) F 0 , X Ω . ( 8 - 10 )

where F0 is the macroscopic deformation corresponding to the boundary conditions of the MVE, F(X) is the local deformation at point X in the MVE domain Ω. In two dimensions, A(X) has (2×2)2=16 independent components, requiring direct numerical simulations (DNS) under four orthogonal loading conditions to uniquely define. In three dimensions, A(X) has (3×3)2=81 independent components, requiring DNS under nine orthogonal loading conditions to uniquely define. However, for specific problems where the loading condition is known a priori, the deformation gradient (or strain) field of the same loading condition can be used for clustering. Once the clustering data is prepared, clustering methods, such as k-means clustering or self-organizing maps, can be used to find a predefined number of clusters.

The interaction tensor has to be recalculated every time C0 is updated during the self-consistent scheme. However, most of the calculation effort can be done in the offline stage if C0 is isotropic. An isotropic reference stiffness tensor can be expressed as


Cklmn00δklδmn+2μ0δkmδln,  (8-11)

where μ0 and λ0 are reference Lamé constants. The corresponding Green's operator Γ0 can be decomposed into two parts:


Γ0=c1001+c2002.  (8-12)

Here, c1 and c2 depend on λ0 and μ0:

( c 1 = 1 2 μ 0 , and c 2 = - λ 0 2 μ 0 ( λ 0 + 2 μ 0 ) . ( 8 - 13 )

The terms Γ1 and Γ2 have simple forms in the Fourier frequency space:

( Γ ^ ijkl 1 = δ ik ξ j ξ l ξ 2 , and Γ ^ ijkl 2 = ξ i ξ j ξ k ξ l ξ 4 , ( 8 - 14 )

where ξ is a Fourier frequency point and |ξ|=√{square root over (ξiξi)}. Thus, the interaction tensor can be expressed as


DIJ=c1D1IJ+c2D2IJ.  (8-15)

where

D w IJ = 1 Ω I Ω χ I ( Γ w * χ J ) d Ω , w = 1 , 2. ( 8 - 16 )

Notice in Eqs. (8-14) and (8-16) that DwIJ, w=1,2 do not depend on the two parameters λ0 and μ0, thus need only be calculated once, which is done in the offline stage. If the MVE can be represented by a regular grid (i.e., voxels), the convolution in Eq. (8-16) can be obtained with relatively little computational effort using an FFT algorithm:


ΓwJ=−1({circumflex over (Γ)}wJ)),  (8-17)

where is the FFT operation and −1 is its inverse.

Procedure for Polycrystalline Microstructure-Property Database Generation

In certain embodiments, the implementation procedure for generating a polycrystalline microstructure-property database includes the following steps.

    • 1. If not using images, set microstructure descriptors; else
      • (a) load 3D images;
      • (b) measure microstructure descriptors;
    • 2. Run Dream.3D pipelines to generate M MVEs;
    • 3. Initialize m=1;
    • 4. Set number of clusters per grain k for MVE m;
    • 5. For MVE m, run CPSCA offline calculations:
      • (a) if k=1, go to (d); else continue;
      • (b) elastic strain concentration calculation using FFT;
      • (c) domain decomposition using k-means clustering;
      • (d) interaction tensor calculation using FFT;
    • 6. Set loading conditions;
    • 7. Run CPSCA online subroutine;
    • 8. Evaluate and add effective properties to database, and m←m+; and
    • 9. Repeat steps 4-8 until m=M.

Online Stage: Self-Consistent Scheme

For large deformation, the far field deformation gradient is applied incrementally. The incremental far field deformation gradient ΔF0 of the current loading step is defined by Fcurrent0−Flast0. Then an incremental form of Eq. (8) is given by


ΔFIJ=1NcDIJ:(ΔPJ−C0:ΔFJ)−ΔF0=0, for I=1, . . . ,Nc,  (8-18)

where ΔFJ and ΔPJ are the local incremental deformation gradient and PK1 stress. Since ΔPJ can be determined as a function of ΔFJ through a local constitutive law in the Jth cluster, the unknows for Eq. (8-18) are {ΔF}={ΔF1, ΔF2, . . . , ΔFNc}. The residual form of Eq. (8-18) given in Eq. (8-19) can then be solved in the online stage, using Newton's iterative method.


rI=ΔFIJ=1NcDIJ:(ΔPJ−C0:ΔFJ)−ΔF0, for I=1, . . . ,Nc  (8-19)

The system Jacobian {M} is defined component-wise as:

M IJ = r I Δ F J = δ IJ I 4 + D IJ : ( C J - C 0 ) , for I , J = 1 , , N c , ( 8 - 20 )

where

C J = Δ P J Δ F J

is the tangent stillness tensor of the material in the Jth cluster and is an output of the local constitutive law in the cluster for the current loading increment. I4 is a 4th rank identity tensor defined by I4,klmnkmδln, and δIJ is the Kronecker delta in terms of indices I and J. The Newton's method update for the incremental deformation gradient can then be expressed as


F}=−{M}−1{r}.  (8-21)

As mentioned, a self-consistent iterative scheme is also necessary to update the isotropic C0 at each loading increment so that it approximates the effective tangent stiffness Ceff of the MVE at that loading step. The general Ceff can be obtained by noting that

C e f f = Δ P 0 Δ F 0 = I = 1 N c v I Δ P I Δ F 0 = I = 1 N c v I Δ P I Δ F I : Δ F I Δ F 0 = J = 1 N c v I C I : A I , ( 8 - 22 )

where ΔP0 is the incremental far field PK1 stress of the current loading step given by ΔP0I=1Nc vIΔPI; vI is the volume fraction of cluster I

A I = Δ F I Δ F 0

is the deformation gradient concentration tensor of the Ith cluster for the current loading increment. For each increment, this is computed by noting that the differential form of Eq. (18) is d{ΔF0}={M}d{ΔF}, which gives d{ΔF}={M}−1d{ΔF0}, where {ΔF0}={ΔF0; . . . ; ΔF0} has Nc blocks of ΔF0. Denote BIJ as the IJ component of the inverse of the Jacobian system: {B}={M}−1, then AIJ=1Nc BIJ.

The isotropic C0 is obtained by minimizing ∥ΔP0−C0:ΔF02, where ∥Z∥2=Z:Z for an arbitrary second-order tensor Z. The drawback of this method is that the optimization problem is under-determined under pure shear or hydrostatic loading conditions. In this exemplary example, the approximation is done by projecting the Ceff of the MVE to a 4th rank isotropic tensor. This is done by first expressing the isotropic C0 as


C0=(3λ0+2μ0)J+0K,  (8-23)

where the forth-rank tensors J and K are defined as


J=I2⊗I2 and K=I4−J.  (8-24)

Then by noting that


J::K=0,J::J=1, and K::K=8,  (8-25)

the two Lamé constants are obtained as

( λ 0 = 1 3 ( C eff : : J - 1 8 C eff : : K ) μ 0 = 1 16 C eff : : K ( 8 - 26 )

where :: defines the double contraction operation.

Algorithm for the Online Stage of SCA

In certain embodiments, the algorithm for the online stage of SCA includes the following steps.

    • 1. Initial conditions and initialization:
      • (a) set n=0, {F}n={I2}, {P}n=0, {ΔF}n=0 and {ΔF}new={ΔF}n;
      • (b) call a UMAT subroutine to get CJ, J=1, . . . , Nc, and set the reference stiffness Cn0J=1NcvJCJ;
    • 2. For load increment n+1, update the interaction tensor parts {D1} and {D2};
    • 3. Newton iterations:
      • (a) call the UMAT subroutine to get {ΔP}new and {C};
      • (b) compute the residual {r};
      • (c) compute the system Jacobian {M}=∂{r}/∂{ΔF};
      • (d) solve the linear equation {M}{δF}=−{r} for {δF};
      • (e) {ΔF}new←{ΔF}new+{γF};
      • (f) if maxJ=1Nc{∥δFJ∥}<tolnewton is not met, go to 3(a);
    • 4. Calculate effective tangent stiffness Ceff and project it to a 4th rank isotropic tensor to obtain Cn+10;
    • 5. If ∥Cn+10−Cn0∥<tolsc is not met, go to step 2;
    • 6. {F}n+1←{F}n+{ΔF}new, {P}n←{P}n+{ΔP}new, n←n+1 and update the state variables for the UMAT subroutine;
    • 7. Repeat steps 2-6 until the simulation completes.

Numerical Validation

By formulating the MVE problem at finite strains, we are now able to correctly consider general elasto-viscoplastic constitutive laws of the type that rely on the multiplicative decomposition of the deformation gradient. One common application for such a law, commonly known as crystal plasticity, which describes the mechanical behavior of a single crystal, is given. The macroscopic response of a polycrystal aggregate predicted by SCA with such a material law is validated by comparing with two full-field methods. It is shown that SCA achieves comparable results with significantly reduced DoFs and computational expense.

Elasto-Viscoplastic Material Model

In the general elasto-viscoplastic material model, the local deformation gradient F is multiplicatively decomposed into elastic Fe and inelastic Fin contributions:


F=Fe·Fin.  (8-27)

Fe is a combination of the elastic stretch and rigid body rotation, while Fin is associated with unrecoverable deformation mechanisms, such as dislocation slip and/or transformation plasticity. The elastic constitutive law is given by


Se=CSE:EeCSE:[(Fe)T·Fe−I2],  (8-28)

where Ee is the elastic Green-Lagrange strain, Se is the Second Piola-Kirchhoff stress, CSE is the 4th order elastic stiffness tensor, and I2 is the 2nd order identity. Taking the inelastic term as solely the plastic deformation (Fin=Fp), the inelastic deformation gradient Fin can be determined using a plastic constitutive law to relate the plastic velocity gradient Lp={dot over (F)}p·(Fp)−1 to plastic shear rate {dot over (γ)}α in slip system a through


Lpα=1Nslip{dot over (γ)}α(s0α⊗n0α).  (8-29)

Here, s0α and n0α are unit vectors which define the slip direction and slip plane normal for slip system α in the undeformed configuration, Nslip is the number of active slip systems, and ⊗ is the dyadic product. In general, the plastic shear rate {dot over (γ)}α in slip system a is taken to be a function of resolved shear stress τα, deformation resistance τ0α, and back stress aα in that slip system. The resolved shear stress is given by


τ(α)=σ:(sα⊗nα),  (8-30)

where σ is the Cauchy stress, s(α) is the slip direction, and n(α) is the slip plane normal, all of which are defined in the deformed configuration. They are computed from their counterparts in the undeformed configuration with

( σ = 1 J e [ F e · S e · ( F e ) T ] , s α = F e · s 0 α , n α = n 0 α · ( F e ) - 1 . ( 8 - 31 )

It is possible to use any appropriate evolution law for {dot over (γ)}α. In this work we choose to employ a power law for {dot over (γ)}α given by

γ . a = γ . 0 τ α - a α τ 0 a ( m ~ - 1 ) ( τ α - a α τ 0 α ) , ( 8 - 32 )

where {dot over (γ)}0 is a reference shear rate, and {tilde over (m)} is the exponent related to material strain rate sensitivity. The evolution laws for deformation resistance τ0α (the isotropic hardening term) and back stress aα (the kinematic hardening term) are given:

( τ . 0 α = H β = 1 N slip γ . β - R τ 0 α β = 1 N slip γ . β , a . α = h γ . α - r a γ . α ( 8 - 33 )

where H and h are direct hardening coefficients, and R and r are dynamic recovery coefficients. Note that in Eq. (8-33) we assume the latent hardening and self-hardening effects are identical. To account for the grain size effect on apparent properties, a Hall-Petch-type equation is introduced that relates the initial slip system deformation resistance in a grain, τ0α,t=0, to the equivalent sphere diameter (ESD) D of that grain with

τ 0 α , t = 0 = τ 0 , in α , t = 0 + K α D , ( 8 - 34 )

where τ0,inα,t=0 denotes the intrinsic initial slip resistance, and Kα is the grain size strengthening coefficient. This equation approximates the impeding effect of grain boundaries on dislocation slip.

Given a deformation gradient increment ΔF, its corresponding increment of PK1 stress can be calculated following the numerical algorithm given. 9 also provides the tangent stiffness

Δ P Δ F

used to couple this material law with SCA.

Tensile Behavior of a Grain Aggregate

The material considered in this work is a fully transformed α-phase Titanium alloy, containing 24 active hexagonal close packed (HCP) slip systems: 3 <1120>{0001} basal, 3<1120>{1010} prismatic, 6<1120>{1011} first order pyramidal and 12<1123>{1011} second order pyramidal. The elastic stiffness matrix is assumed to be transversely isotropic with components given in Table 8-1, where Cij are the stiffness components in Voigt notation. The material parameters for the plasticity law used here are given in Table 8-2; the first four entries use the calibrated values, and the remaining entries are assumed.

TABLE 8-1 Transversely isotropic elastic stiffness components for a HCP titanium alloy Component Value (MPa) Component Value (MPa) C11 = C22 1.70 × 105 C33 2.04 × 105 C44 = C55 1.02 × 105 C66 0.36 × 105 C12 = C21 0.98 × 105 C13 = C31 0.86 × 105 C23 = C32 0.86 × 105 other Cij 0

TABLE 8-2 Crystal plasticity parameters for a HCP titanium alloy. Pyramidal Basal Prismatic Pyramidal <a> <c + a> {dot over (γ)}0α (s−1) * 0.0023 0.0023 0.0023 0.0023 {tilde over (m)}[*] 50 50 50 50 τ0, inα, t=0 (MPa) * 284.00 282.24 395.00 623.30 Kα (MPa√{square root over (mm)}) * 16.45 16.45 16.45 16.45 αα, t=0 (MPa) 0.0 0.0 0.0 0.0 H (MPa) 1.0 1.0 1.0 1.0 R (MPa) 0.0 0.0 0.0 0.0 h (MPa) 500.0 500.0 500.0 500.0 r (MPa) 0.0 0.0 0.0 0.0 * J Thomas, M Groeber, and S Ghosh. Image-based crystal plasticity FE framework for microstructure dependent properties of Ti—6Al—4V alloys. Materials Science and Engineering: A, 553: 164-175, 2012.

The microstructure considered is an idealized MVE of eight equal-sized 40 μm×40 μm×40 μm cubic grains, as shown in panel (a) of FIG. 109 as colored according to the inverse pole FIG. 8—given in panel (b) of FIG. 109, which is defined with a z-face normal as the reference direction. This allows both the FEM and FFT to use the same mesh and mesh refinement. The MVE is subjected to uniaxial tensile deformation in the x-direction with a strain rate of 10−4/sec until a maximum strain of 0.02. This deformation is applied in 200 increments using three methods: SCA, FEM, and FFT; hereafter these will be termed CPSCA, CPFEM and CPFFT where CP is short for crystal plasticity. For CPFEM, Abaqus/Standard is used with C3D8R elements and a user material implemented as a UMAT for implicit analysis. For CPFFT, the Newton-Krylov solver is implemented. For CPSCA, the offline clustering data is prepared with the material modeled as anisotropic-elastic and with the same overall loading history. The elastic strain field at peak remote strain is used as the offline data to obtain clusters using k-means clustering. Panel (c) of FIG. 109 shows the distribution of the elastic strain component E11 of the offline data obtained using FEM with a 10×10×10 voxel mesh. Panel (d) of FIG. 109 shows the distribution of 128 clusters (16 clusters/grain) obtained using the k-means clustering method.

To show the accuracy of CPSCA, the converged macroscale stress-strain curve from full-field methods is used as the reference solution. The converged solution is obtained by refining the mesh to 60×60×60, which results in 648,000 DoFs for CPFEM and 1,944,000 for CPFFT. Panels (a)-(b) of FIG. 110 show that the elastic response is not sensitive to mesh refinement, while a substantial difference exists in the plastic deformation regime between the results of CPFEM and CPFFT with a course mesh, and quite similar results are achieved with a refined mesh. In the following comparison, the reference solution is the CPFFT result with a 60×60×60 grid, or 27,000 grid points/grain. Four different number of clusters per grain, ranging from 1 to 64 (total DoFs ranging from 72 to 4608), are used to show the convergence of CPSCA. Panels (c)-(d) of FIG. 110 show that the CPSCA solution is indistinguishable (within 1% difference) even with only 1 cluster/grain in the elastic deformation regime. In the plastic deformation regime, the CPSCA solution approaches the CPFFT solution as an increasing number of clusters is used.

CPSCA achieves high efficiency by solving only a few DoFs, without sacrificing much accuracy. To demonstrate this, the 0.2% plastic strain offset (σ0.2, near the yield point) and 0.4% offsets stresses (σ0.4, away from the yield point) are compared to the reference solution for different number of clusters. This difference is plotted in panel (a) of FIG. 111. The difference between σ0.2 s as a function of number of clusters oscillates, but is always below 0.4%. Conversely, the difference between σ0,4 s can be as high as 1.4%, but decreases monotonically as the number of clusters increases. Panel (b) of FIG. 111 shows that the CPU time in both offline and online stages required for CPSCA increases as a function of the number of clusters/grain used. The slope on logarithmic axes is around two for the offline stage and around three for the online stage. This is because in the offline stage the number of interaction tensors among the clusters that need to be computed is Nc2. In the online stage, the computational complexity of matrix inversion is O(Nc3). Using 64 clusters/grain (amounting to 512 total clusters) takes comparable time to using CPFFT with a 40×40×40 grid, but is two orders-of-magnitude faster than CPFEM using the same mesh. The increase in computational time with increasing number of clusters is driven mostly by the online stage if a large number of clusters are used.

Case Study 1: Uncertainty of Predicted Effective Properties

One application of microstructure modeling is to evaluate the effective properties of a bulk material. For example, under monotonic tensile loading, a virtual tension test predicts the overall stress-strain curve directly through homogenization. From this, common scalar material properties such as effective elastic stiffness and yield strength can be extracted. Under cyclic loading, local stress and strain information can be used to evaluate fatigue life. Due to restrictions on the MVE size because of computational expense and microstructure randomness, uncertainty quantification of the predicted effective properties is needed. To achieve this, multiple realizations of the MVE homogenization are computed using the SCA method introduced above. A flowchart for generating a MVE-property database is shown in FIG. 112, which is implemented with the algorithm. In this section, this procedure is used to quantify the uncertainty of yield strength under different textures and the grain size effect.

Synthetic Microstructure Volume Elements

Although MVEs can be obtained by extracting volumetric information from 3D experimental images or grain growth simulations, as mentioned in the introduction, such an approach could be very expensive because many MVEs are needed to generate a microstructure-property database. More practically, a statistical distribution of microstructure feature parameters can be measured from routine, 2D experimental characterization and used to synthetically construct MVEs. To construct polycrystalline MVEs, we use the software package DREAM.3D, which includes tools to generate microstructures that adhere as nearly as possible to predetermined statistics of descriptors. This work focuses on varying two microstructure descriptors of polycrystalline materials: grain orientation distribution function and grain size distribution function.

The crystallographic orientation distribution function (ODF), also known as texture, defines the probability density ƒ(Q) of crystallites falling into an infinitesimal neighborhood around the orientation Q, which is often parameterized by Euler angles. In this work, we use the Bunge convention of Euler angles (ϕ1, Φ, ϕ2), which define subsequent rotations about the z-axis, then the new x-axis, and then the new z-axis again. In DREAM.3D, the orientation space is discretized into bins and an ODF with strong texture can be generated by specifying some orientations, the corresponding weights defined as multiples of random distribution (MRD) and number of bins it takes for the MRD value to reduce to zero decreasing quadratically from the bin of the entered orientation.

The grain size distribution function is assumed to be log-normal with the probability density function given by

f ( D ) = 1 D · 1 σ 2 π exp ( - ( ln D - μ ) 2 2 σ 2 ) , (8-35)

where D is the equivalent sphere diameter (ESD) of a grain, σ a scale parameter, and μ is the shape parameter. Other descriptors such as misorientation distribution function, aspect ratio, and number of neighbors are also necessary to synthesize a complete microstructure realistically

Effect of Texture on Effective Properties

To show the effect of texture on yield strength, four texture cases were considered: no texture, (0,0,0) preferred, (90,45,0) preferred, and (90,90,0) preferred. Fifty cubic MVEs were synthetically generated using DREAM.3D with grain size distribution parameters μ=19.7 μm and σ=2.7 μm for each of the four texture cases. Each MVE has around 90 equiaxed grains represented in a 81×81×81 voxel mesh with a resolution of 1 μm×1 μm×1 μm per voxel. Example MVEs for these texture cases are given in FIG. 113. All MVEs were subjected to monotonic uniaxial tension in the x-direction with strain rate 10−4/sec until a maximum strain of 0.02. The stress-strain curves predicted using CPSCA with one cluster/grain are given in FIG. 114. The effective Young's modulus and 0.2% offset yield strength are measured from each of these predictions and compared in FIG. 115. MVEs with preferred (90,90,0) texture have the highest yield strength on average. MVEs with preferred (90,0,0) texture and (90,45,0) texture have approximately the same yield strength on average and they are the lowest. MVEs without texture are in between. The variations in yield strength for each case are also shown.

Effect of Grain Size on Effective Properties

The effect of grain size on yield strength is studied by setting σ in Eq. (8-35) such that four different average grain sizes are generated between 10 μm and 40 μm. Fifty MVEs of equiaxed grains without texture are generated using DREAM.3D for each σ value while keeping all other parameters the same. Each MVE is represented in a 81×81×81 voxel mesh. Sample MVEs with different grain size are given in FIG. 116. All MVEs are subjected to monotonic uniaxial tension in the x-direction with strain rate 10−4/s until a maximum strain of 0.02. FIG. 117 shows the distribution plots of predicted effective Young's modulus and 0.2% offset yield strength versus averaged ESD. The average Young's modulus is not sensitive to average ESD. A Hall-Petch type relationship between the average yield strength and average ESD is also observed as expected. However, the uncertainty increases with the average ESD as the number of grains in the MVE decreases. This is because fewer grains makes the MVE less representative.

Case Study 2: Concurrent Multiscale Simulation of a Rolling Process

Another type of application for high-efficiency, microstructure-based modeling is the simulation of finite-deformation processes, in which microstructure evolves extensively and gives rise to complex macroscopic behaviors. For such applications, it is often hard to find a simple-form phenomenological constitutive model at the macroscale. Moreover, a phenomenological constitutive model has to be recalibrated for new materials in which microstructures are different. In this case study, it will be shown that the speed of the CPSCA-based microstructure modeling makes possible a concurrent multiscale model for applications where finite strains at the microscale are crucial. Our exemplar is the metal rolling process. Texture evolution is important at the microscale, and it depends on the macroscale loading and distribution of stress throughout the part. To capture all this, a multiscale method is required.

A schematic of the rolling process of a thick plate is shown in panel (a) of FIG. 118. The initial thick plate has height (H) of 40 mm, thickness (T) of 40 mm and length (L) of 92 mm. The radius of the roller is 170 mm with a rotation speed of 2 m/second resulting a rolling speed (roller surface speed) of 1.07 m/second. The height of the rolled plate is assumed to be 30 mm. In the initial thick plate, the microstructure is assumed to be equiaxed HCP grains with random lattice orientations (no texture) and average grain size of 26 μm. The simulation goal is to predict part deformation as well as microstructure evolution during the rolling process. Direct numerical simulation of this system, by modeling all grains explicitly (more than 20 billion grains for the thick plate studied here), exceeds current modern computational capability. An alternative way is to use the two-scale concurrent multiscale simulation method schematically shown in panel (b) of FIG. 118. The plate (macroscale) is fully coupled with a MVE (microscale) in such a way that the deformation gradient of a macroscale integration point is passed to its associated MVE as far field loading, then the MVE problem is solved and the homogenized stress is passed back to the macroscale integration point. Most studies in the literature using this approach with full-field methods e.g., FE2 for the microscale MVE problem are computationally limited to 2D problems. We will show in this case study that CPSCA enables realistic 3D simulation of microstructure evolution with reasonable computation time.

For the macroscale problem, the implicit time integration method was employed so that larger time increments can be used. Coulomb friction was assumed between the roller and the plate, with a friction coefficient of 0.3 for the plate being pulled through the roll stand. The thick plate was given an initial velocity of 1.07 m/second to reduce impact between the plate and the roller which might cause numerical difficulty. By taking advantage of symmetry, only 1/4 plate and a single roller were modeled. The 1/4 plate was meshed with 2994 uniform hexahedral elements and the roller was modeled with fine enough rigid elements. Reduced integration was used with stiffness-based hour-glassing control. For the microscale problem, the MVE was chosen to include 90 randomly orientated equiaxed grains (panel (b) of FIG. 118) and each was represented by one cluster. The macroscale dynamic analysis was implemented with the commercial software Abaqus/Standard and the microscale CPSCA method was implemented as a user material subroutine, or UMAT. 8 provides the steps to calculate the stress at the end of a time step in a UMAT. The tangent stiffness tensor required by Abaqus/Standard is calculated with Eq. (8-22).

The simulation was stopped at rolling time 0.1 seconds. Panels (a)-(b) of FIG. 119 show the contours of shear stress σ12 after 0.08 seconds of rolling, predicted with the multiscale 3D simulation. For comparison, the 2D plane strain simulation result is shown in panel (c) of FIG. 119. This shows that on the symmetric plane, the 3D simulation and 2D simulation give similar shear stress distribution patterns: the alternating positive and negative shear stress value near the contact region between the plate and the roller. However, the 3D simulation shows that shear stress is not uniform in the thickness direction (z-direction). The 3D simulation predicts lower deformation in the rolling direction and higher extreme shear stress. The variation of σ12 as rolling progresses for the 3 elements indicated by the red arrows in panel (a) of FIG. 119 is given in FIG. 120. For all 3 elements, σ12 alternates and reaches a stable, nonzero value in the end. The shear stress value of elements closer to the contact region tend to alternate more times and with higher amplitude.

Another advantage of the concurrent multiscale simulation is that microstructure evolution is solved for the whole manufacturing process. FIG. 121 shows snapshots of MVE deformation and (0001) pole figures (generated with MTEX) for each integration point of the three elements during the rolling process at the time points indicated by the vertical dashed red lines in FIG. 120. This shows that MVEs with different position in the macro part relative to the rolled face deform and rotate differently: away from the contacting surface (e.g., compare Element A to Element C), there is less crystallographic rotation and shear, and more compression. The rotation of each grain can be seen from the pole figures, where each dot indicates the lattice orientation of a grain plotted on a plane using the stereographic projection. During rolling, these dots concentrate around the two ends in the lateral direction of the circle meaning that the grain orientation rotates towards the y-direction (the direction of maximum compression). This matches the experimental observation of the texture of cold rolled, pure α-phase Titanium. Such texture is deemed to be caused by dislocation slip which is captured by the current model. Note that texture due to twining is different. In order to predict the evolution of the Ti system more accurately, our model could be extended to capture twinning.

By using the Abaqus MPI parallelization with 72 cores (on three nodes each with two Intel Haswell E5-2680v32.5 GHz 12-core processors), the concurrent simulation takes approximately 112 h.

In sum, a finite-strain self-consistent clustering analysis is developed and applied to model interacting and evolving microstructure for general elasto-viscoplastic materials. The method is reformulated in an initial Lagrange configuration so that large deformation problems can be solved. The accuracy and efficiency obtained for the prediction of overall mechanical response of polycrystalline materials is demonstrated by comparing with both the finite element analysis and the fast Fourier transform-based method. It is shown that CPSCA achieves high accuracy with significantly reduced degrees of freedom.

In our case studies, grains are resolved explicitly in voxel-based MVEs reconstructed with predefined microstructure descriptors. These MVEs are used to predict with quantitative uncertainty the influence of texture on yield strength, and grain size on yield strength. Finally, a concurrent multiscale simulation of a rolling process shows the heterogeneous microstructure evolution throughout the rolled part. This is made possible by the efficiency of this method. Potential applications include simulation-driven microstructural design and manufacturing process control.

Example 9 Modeling and Characterization of Integrated Computational Materials Engineering (ICME) Composites 1. Atomistically Informed Resin Infusion Model

In this work, predictive atomistic models of epoxy resins are developed, and the thermomechanical properties and their dependence on the molecular chemistries of the resin matrix are characterized, including resin functionality, crosslink degree, and component ratio, demonstrating the viability of utilizing atomistic simulation to predict key material properties and trends. In addition, we also presented a hierarchical multiscale model where MD simulation results were homogenized to a thermo-plastic law to describe the constitutive behavior of epoxy resins. This thermo-plastic law has been used in RVE modeling to predict the stiffness and strength of CFRP composites. Furthermore, we characterized the properties of the nanoscale interphase between carbon fibers and resin matrix and integrated the interphase into the mechanistic continuum models for CFRP, and elucidated the explicit effect of the interphase region on the failure behavior of the composites, which generated insights to guide future design strategies for failure-resistant composites.

The superior thermomechanical properties of epoxy resins have led to a wide range of applications, most notably as matrix materials in fiber-reinforced composites. The excellent thermomechanical properties arise from the highly crosslinked molecular structure the resins could form. Nanoscale simulations of epoxy resins offer a promising way to characterize their properties and the dependence on molecular-level factors, such as resin type and crosslink density. Furthermore, a deep understanding of the dependence of thermomechanical properties on the molecular-level structures is of critical importance to guide future computation-based design for epoxy resins with optimized mechanical properties.

Atomistic molecular dynamics (MD) simulations on epoxy resins have been successfully applied to predict various material properties. Several computational algorithms have been developed to generate reasonable crosslinked structures for investigation of their physical properties. MD simulations have been carried out to predict the glass transition temperature (Tg) and provided valuable insights into the effects of strain rate, temperature, and crosslink degree on Young's modulus and yielding behavior. Despite significant progress toward understanding epoxy thermomechanical response, multiscale models that can bridge length and time scales, especially couple atomistic and continuum scales, remains a particular challenge.

To overcome this challenge, we first developed nanoscale models of representative epoxy resins by capturing the specific crosslinked structures. We then characterized elastic, yield, and post-yield behavior from MD simulations. After that, yield surfaces were generated from MD simulation results, which can be well described by a paraboloid yield criterion. Further, by adding plastic potential and hardening law, a thermo-plastic law was proposed to describe the constitutive behavior of epoxy resins. Along the way we also illustrated the dependence of thermomechanical properties of epoxy resins on molecular chemistry, such as epoxy type, component ratio, and crosslink density.

In addition, the interphase region that exists between fibers and resin matrices possesses heterogeneous chemical and physical features and has a thickness at the sub-micron scale. Despite being much smaller than the fiber diameter, the interphase region has been shown to play a critical role in the performance of CFRP composites. Accurate modeling or characterization for the interphase region remains a significant challenge. To overcome the hurdles encountered in nanoscopic experiments, efforts have been reported to characterize the interphase region using analytical models or MD analyses. However, there have been few studies that integrate the nanoscale interphase region with RVE modeling and study the effect of the interphase region on the macroscopic composite response. To address this issue, we first obtained the properties of the interphase region according to MD simulation results and a generic analytical gradient model. Then, the average property of the interphase region was incorporated into a modified RVE model, in which the three phases, fiber, matrix, and the interphase, were included. This modified RVE model was shown to improve significantly in predictions of the modulus and failure strength of the composites.

Generating the Realistic Crosslinked Structure of Epoxy Resins and Yield Surface Calculation

We chose two representative epoxy systems as our model system: (1) an epoxy resin commercially known as Epon 825, including diglycidyl ether of Bisphenol A (DGEBA) with curing agent 3,3-diaminodiphenyl sulfone (33DDS); and (2) an epoxy commercially denominated as 3501-6, mainly composed by tetraglycidyl methylenedianiline (TGMDA) with curing agent 4,4-diaminodiphenyl sulfone (44DDS). We integrated the Polymatic Algorithm with the LAMMPS package to simulate the crosslinking process. Basically, covalent crosslink bonds were added between eligible atoms based on pair-wise separation distance. Also, for every several crosslink bonds formed, energy minimization and equilibration simulations were conducted with MD to alleviate the stress generated. This workflow was able to generate atomistic structures of epoxy resins with different crosslink degrees from different initial chemistries and component ratios.

To obtain the yield surface of typical epoxy resins, the stress-strain responses of the Epon 825 model system were first calculated from the MD simulations at different temperatures and at a strain rate of 5×108 s−1. We note that the high strain rate is inherent in MD simulations given the small time-step used. During these simulations, proper thermostatting is applied to maintain the systems at specified temperatures. The results for uniaxial tensile and compressive loading cases are plotted in FIG. 122. As can be seen in the figures, the entire stress-strain response for both loads is temperature dependent, affecting the yield stresses and the elastic moduli. This behavior is well-known in MD simulations of glassy polymers.

The subsequently obtained yield surfaces for the model system at different temperatures is shown in FIG. 123. We adopted the common convention where yield stress corresponds to the maximum point or at the obvious “knee” in the stress-strain curve. We find that there is a good agreement between the MD results with the paraboloidal yield surface. This yield criterion is determined uniquely by two material parameters, the compressive and tensile yield stresses:


ƒ(σ,σγcγT)=6J2+2(σc−σT)I1−2σcσT  (9-1)

where J2 is the second invariant of the deviatoric stress tensor, and I1 is the first invariant of the stress tensor. σT and σc denote the tensile and compressive yielding stress, respectively.

Due to the high strain rates at which MD simulations are performed, the yield stresses obtained are higher than the values obtained experimentally. Nevertheless, we further find that the experimental results on yield stresses can be well described by the same criterion as shown in Eq. (9-1).

FIG. 123 shows yield surfaces obtained for different temperatures where the points are simulation data and the lines are theoretical prediction using Eq. (9-1).

Similar to other plasticity formulations, the thermo-plastic law disclosed herein is then defined by the yield surface, plastic potential, and hardening law as outlined next.

First, a plastic potential with a non-associative flow rule such that a positive volumetric plastic strain is prevented under hydrostatic pressure is defined as:


g=σvm2+αp2  (9-2)

where σvm=√{square root over (3J2)} is the von Mises equivalent stress, P=1/3 I1 is the hydrostatic pressure, and α is the material parameter to correct the volumetric component of the plastic flow, which equals to

α = 9 2 1 - 2 v p 1 + v p (9-3)

with vp being the plastic Poisson's ratio. This thermo-plastic law of the resin matrix has been integrated into the mechanistic continuum models for CFRP with basic parameters informed by experimental results. By using this law, the characterized yield surfaces for the epoxy resins agree very well with experimental results, with the error less than 5%, thus achieving the goal and objective of the project.

The framework for developing crosslinked epoxy resin structures as well as yield surface characterizations is generally applicable to other epoxy resin systems with different chemistries.

Dependence of Thermomechanical Properties on Molecular Chemistry

We have also studied the large-deformation behavior of epoxy resins and characterized their failure response at the atomistic level. During large deformation, there are inevitable bond breaking events happening in the network structures of epoxy resins. To capture the realistic bond breaking phenomena, we adopted a reactive force field, which has been validated to preserve the elastic and plastic responses of the epoxy resins studied here. Stress-strain curves of 3501-6 epoxy systems with different crosslink degrees and component ratios are plotted in FIG. 124. Consistent “elastic-yielding-hardening-failure” behavior is observed for all the cases. With increasing crosslink degree, both yield and maximum stresses increase, which is associated with decreasing failure strain or deformability. Varying the component ratio has a subtler change in the stress-strain curves, but the stoichiometric one has the highest yield stress and maximum stress while the lowest deformability. Thus, from atomistic level tensile simulations, we showed that the molecular-chemistries of resins strongly influence their mechanical properties and failure responses.

Building upon the stress-strain curves from tensile simulations and the parameters quantifying the structural changes such as chain reorientation and void formation, we have linked this atomistic level failure response of resins to their macroscopic fracture properties on the basis of a continuum fracture mechanics model. This work provided physical insights into the molecular mechanisms that govern the fracture characteristics of epoxy resins and demonstrated the success of utilizing atomistic simulations toward predicting macroscopic fracture energies.

We would like to note that the planned methodology to investigate the large-deformation and failure behavior of epoxy resins was to develop coarse-grained models for epoxy resins, which could increase the computational efficiency. We departed from the planned methodology by using atomistic simulations with specific chemistry details captured. Although being more computationally expensive, atomistic simulations gave us direct predictions of the thermomechanical properties while avoiding the need to calibrate force fields for coarse-grained models. Additionally, reactive force field provides more accurate predictions for the stress-induced bond breaking events and failure responses of the resins at the nanoscale. More importantly, the multiscale methodology by informing yield surface criterion and thermo-plastic constitutive laws is more powerful to bridge length and time scales than coarse-grained models.

Interphase Property Characterization

Due to the surface roughness of carbon fibers, the surface treatments during fiber manufacturing process, and matrix affected regions, there exists a submicron-thick interphase region around carbon fibers. The thickness of the interphase region has been evaluated to be about 200 nm with an analysis from transmission electron microscopy (TEM). Here, the interphase region is further simplified as a cylindrical shell adjacent to the fiber, with the inner radius rf being the same as the fiber radius and outer radius ri=rf+200 nm, as shown in FIG. 125. In the following text, sub-indices ƒ, i and m denote fiber, interphase region and matrix, respectively. Although there has not been a quantitative characterization of the interphase region in situ, we know some basic information of the property variation inside the interphase region. First, at the inner and outer boundaries, both physical and chemical properties of the interphase comply with the adjacent phases. Second, there exists a sharp gradient from fiber property to the matrix property within the interphase region. Third, due to the incompatibility between the sizing of the fiber surface and the resin matrix, we anticipate part of the regions within the interphase achieve a lower crosslink degree. The experimental observation that the failure initiates inside the interphase region provides further evidence of this weak region. In the previous section and FIG. 124, we have characterized the effect of crosslink degree on the elastic modulus and strength of typical epoxy resins. The results show that the difference of the Young's moduli between insufficiently crosslinked epoxy (about 70% crosslink degree) and fully crosslinked epoxy (95% crosslink degree) is around 20%, and the difference in the strength between them is up to 50%. We use Ems and σms to represent the lower bound values for the Young's modulus and strength inside the interphase region.

To characterize the average properties of the interphase region, we adopted an analytical gradient model to describe the modulus and strength profile inside the interphase. Also, we integrated the MD simulation results on the insufficient crosslinked resins to capture Ems and σms. The gradient model proposed here include two parts. In the first part, Young's modulus and strength decrease from the fiber values to the lowest values, i.e., Ems and σms. In the second part, the values gradually increase from the lowest to the values of the bulk matrix. The decreasing trend in the first part is due to the attenuation of the fiber confinement effect, and the increasing trend in the second part is because of the intrinsic epoxy resin stiffening through sufficient crosslinking. We used the properties of fiber and matrix to formulate the boundary conditions of the interphase region. The position of the lowest values (ris) was assumed to be at three quarters (0.75) of the interphase away from the fiber surface. The position was chosen near the matrix side, since the incompatibility between sizing and bulk matrix resin mainly induces the insufficient crosslinking. A sensitivity analysis has also been conducted to verify that the assumed position of the insufficient crosslink region has a low influence on the average properties of the interphase region.

The variations of the properties of the interphase region were assumed to follow the exponential function as follows:

K i = { K m s + ( K f - K m s ) R ( r ) r f < r < r i s K m + ( K m s - K m ) Q ( r ) r i s < r < r i (9-4)

where K can be either E (Young's modulus) or σ (strength), and the functions R(r) and Q(r) are constructed to match the boundary conditions:

R ( r ) = 1 - ( r / r i s ) exp ( 1 - r / r i s ) 1 - ( r f / r i s ) exp ( 1 - r f / r i s ) (9-5) Q ( r ) = 1 - ( r / r i ) exp ( 1 - r / r i ) 1 - ( r i s / r i ) exp ( 1 - r i s / r i ) ( 9-6 )

The average Young's modulus and strength of the interphase can be obtained as:


KiTfriKi(r)dr/(ri−rf)  (9-7)

Substituting the parameters of both Young's modulus and strength values into the above equation, we finally obtained the average Young's modulus and strength of the interphase. Compared with matrix modulus and tensile strength, the average Young's modulus and strength of the interphase region are increased by around 5 and 9 times, respectively. The interphase region shows an obviously stiffened response compared to the bulk matrix, although a portion of the interphase region is weaker due to insufficient crosslinking. The constitutive behavior and damage model of the interphase were assumed to be similar to those of the bulk matrix. By integrating this stiffened interphase region into RVE model of the UD composites, the accuracy of RVE was much enhanced compared to the traditional two-phase model without the interphase region. The importance of this interphase region is thus clearly manifested.

Our work on atomistic modeling of epoxy resins would provide guidance for future epoxy resin computation-based design. As an important component in integrated computational materials engineering (ICME), atomistic molecular dynamics simulation would further empower the material-by-design process by commercially implementing the technology.

First, nanoscale simulations of highly crosslinked epoxy resins offer a promising way for the development of new continuum theories and models. Fully atomistic models are especially appealing because they are based on fundamental input information—force fields and chemical structure—avoiding the need to calibrate phenomenological laws.

Second, hierarchical multiscale methods which are based on sequential homogenization of smaller scales to larger scales can effectively transfer the information from atomistic or nanoscale to the macroscopic continuum level. In this work, yield surface criterion has been informed from atomistic simulations and integrated into macroscopic models.

Third, the dependence of thermomechanical properties on molecular chemistries revealed here shows promise in accelerating the material-by-design process for thermosets by incorporating data from molecular models. Potential next steps could be leveraging molecular simulations to guide the design of the epoxy chemistry or component ratios in order to optimize the strength and toughness of the thermoset resins.

Last, we have demonstrated that by utilizing molecular simulations and analytical models, we are able to represent the distinct interphase region properties between the fiber and matrix. Subsequently, we have elucidated the explicit effect of this interphase region on the failure behavior of composites. Building upon this, potential future work could involve computational-based design strategies for failure-resistant composites, such as specific nano-engineered architectures and chemistries inside the interphase region.

2. Preform Molding

The preform research uses experimental and computational methods to help understand the material mechanical behavior during the preforming process. Then based on the observation of the material behavior, preforming simulation models with high fidelity are developed. These models start from the macroscopic part-level, progress to the mesoscopic composite level, and finally form a multiscale simulation strategy. The multiscale strategy enables users to have a full understanding of the process parameters optimization, and the lower level composite material design. To validate the simulation models, preforming benchmark tests with shear angle and forming force measurement technique are developed and performed. This benchmark tests, with various combination of process parameters, provide insightful guidance to the preforming process design.

The traditional trial-and-error method to develop a manufacturing process for carbon fiber composites, which relies heavily on experiments, requires great raw material consumption and a long development period. To solve this cost issue, the developed experimental and computational methods for the preforming process form a whole system, which utilizes computation power and virtual manufacturing tools to aid the design and optimization of the carbon fiber composites preforming process.

The experimental research reveals the behaviors of composite prepregs in the manufacturing process, especially the difference from the conventional metal forming process, such as temperature control, fiber reorientation, surface interaction, etc. These behaviors illustrate the importance to adjust the manufacturing technique to the needs of advanced composites. The computational modeling research, on the other hand, completes a whole software package that enables researchers from either academia or industry to virtually design and optimize carbon fiber prepreg preforming, which helps to lower the cost for carbon fiber composite manufacturing development, and broaden the application of this advanced composite material.

To automatically manufacture CFRP parts in large quantities for transportation equipment, thermoforming is a proper choice as it can provide a high production rate with relatively complicated surface geometries, good product quality, and low facility cost. In the thermoforming process, the first step is to stack layers of thermoset carbon fabric impregnated by uncured thermoset resin (prepreg) in an optimized fiber orientation combination. Then, these laminates are heated to soften the resin and subsequently formed to desired 3D shapes on a press machine during the preforming step. Finally, the parts are cured to achieve designed part shapes. In thermoforming, most of the fiber re-orientation is introduced in the preforming step that replaces the conventional high-cost and low-rate hand laying work. Since mechanical stiffness and strength of the composites are mostly affected by the fiber direction, the selection of the preforming parameters such as process temperature and initial fiber orientation is important to the final part performance including shear and kink bands development in the weave under various loading conditions.

To optimize the preforming parameters and produce defect-free parts, numerous tests with different parameter combinations are commonly conducted. However, the consumption of raw material and the long development period increase the cost and time of production; hampering the practicality of thermoforming. To address this issue, several computational models have been developed to simulate the preforming process to predict the fiber orientation, geometry, wrinkling behavior on parts, and forming force. The first widely used computational method to predict the woven CFRP behavior during the preforming process is the pure kinematic-based pin-joint net (PJN) assumption. However, the ignorance of the mechanical properties of the fabric and the resin, results in inaccurate prediction, especially for wrinkling prediction. As an alternative, the finite element method (FEM) draws increasing attention. Simulations for the fiber orientation, draw-in amount and wrinkling behavior prediction during the preforming process have been documented in literatures. Jauffre et al. combined 1D beam elements and 2D shell elements to simulate the tensile and shear behaviors of the material separately. The meshing process for this hybrid element, however, was tricky and time-consuming. Hamila et al. developed a semi-discrete triangle shell element and handled this problem based on internal virtual work. The drawback is that this element was applicable in an in-house FEM software, limiting its usage in the industry. In the LS-DYNA® software, there are built-in woven fabric material models, such as the MAT_234 and MAT 235. Both models, however, are based on mesoscale mechanics and require the input of mesoscale material parameters such as the yarn moduli and yarn-yarn interaction coefficient. It was found in practice that for these parameters, direct experimental characterization is difficult and reverse calculation is time-consuming.

For the potential of commercialization and user-friendly operation, a non-orthogonal material model for the CFRP preforming simulation was developed by Northwestern University and was implemented into a commercial FEM code ABAQUS® as a user-defined material subroutine. Although the intention of coupling the tensile and shear behavior in the new constitutive law was applaudable for having the most general form, it encountered inaccuracy especially when woven CFRP is subject to large shear deformation. As an advancement, an improved non-orthogonal model for the woven CFRP preforming process is invented in the invention. It has been validated by benchmark tests and has been incorporated into the LS-DYNA® software as MAT_COMPRF (MAT_293) through the joint effort of this academic and industry team.

For reliable predictions, there is a need for characterizing and employing realistic and accurate material properties in the computational model. During preforming, prepregs will undergo tension, shear, and bending deformation, as demonstrated in FIG. 126. FIG. 126 also shows that when different prepreg layers have different initial fiber orientation, there will be large relative sliding between prepreg surfaces. As a summary, the targets of experimental characterization for prepreg are: tension, shear, bending, and surface interaction.

The most widely adopted methods for characterizing the properties within one prepreg layer are: 1) the uniaxial tension test to determine the tensile modulus of the composites and 2) the bias-extension test to measure the shear modulus of the woven composites. The reliability and repeatability of these two tests have been validated by using different apparatuses to study a set of similar materials. The bending stiffness of the material is also needed for proper simulation of the material behavior during the forming process. However, the softness of the prepreg under the preforming temperature makes it difficult to measure the bending stiffness via the standard 3-point bending test. An alternative cantilever beam method, which utilizes self-gravity to bend specimens, has been developed and applied to measure the bending stiffness of prepregs. For the prepreg surface interaction characterization, however, a systematic study was still lacking before this integrated computational materials engineering (ICME) project, although this interaction surely affects the fiber orientation after preforming, and plays a significant role in the final product performance. To fill this gap, we designed and built an innovative test apparatus in the invention. Based on our observation of this surface interaction mechanism, a hydro-lubricant interaction model is also constructed to analyze and predict the interaction between prepreg layers. This numerical model together with experiments, reveals the details of the fiber interaction, such as its strength and periodic pattern.

Experimental prepreg characterization can provide reliable results, but it also has some drawbacks. The major disadvantage is that it can only achieve limited loading states. For example, the uniaxial tension test can only introduce pure tension deformation while the bias-extension test can only introduce pure shear deformation. Hence, the coupling between tension and shear cannot be physically characterized and subsequently implemented into the numerical model. Although in most cases this neglection would not affect the prediction of geometry and fiber orientation significantly due to the fact that the shear modulus of the uncured prepreg is always several orders of magnitude smaller than its tensile modulus, it will introduce errors in the prediction of preforming stress and punch force and hence, reduce the analysis accuracy of defects, such as breakage, pull-out, and separation of the fiber yarns.

Several new test devices such as the biaxial tension apparatus and the picture frame apparatus with tension adjustment have been designed to address the above issue. In practice, however, even these complex devices cannot cover all of the strain states that the prepreg will undergo during preforming due to the complexity of three-dimensional geometry and the resulting nonlinear loading paths. Additionally, these experimental characterization methods are at the macroscale and hence do not provide insightful information on how the mesoscale composite structure and constituents affect mechanical properties of the materials. The cost of raw materials and test time also need to be considered in planning experiments. To address this issue, in the invention, we developed a new multiscale preforming simulation method based on the prepreg characterization by the mesoscopic representative volume element (RVE) to account for the tension-shear coupling and applied it to the preforming simulation of a 2×2 twill thermoset prepreg.

Challenges to building a closely packed woven RVE model include structure generation and mesoscopic yarn properties characterization. To construct RVEs with accurate woven patterns and yarn geometrical features, several different approaches have been developed. One approach is to directly use CAD software to design and output the RVE structure. This approach, while being straightforward and suiTable 9—for a specific composite structure, is time-consuming because, for each specific composite, the structure needs to be drawn individually, and when the yarn surface geometries become complex, the yarn cross-section shape needs to be manually identified. To generalize design process and accommodate for more composite structures, geometrical modeling software packages such as TexGen and WiseTex are developed. These packages store large libraries for different composite patterns and can generate the corresponding RVE structure given the geometrical features such as RVE length, yarn width, yarn height, and so on. However, the automatically generated structures usually have a fixed shape of the yarn cross-sections and yarn centerlines. These simplifications are suiTable 9—for loose woven materials but result in yarn-to-yarn penetration in close-packed composites. In this case, fine-tuning the geometry by modifying the position, dimension, or utilizing non-symmetrical shapes of the local yarn cross-section is essential. These procedures, however, involve complicated geometry manipulation and are time-consuming. For capturing more accurate and detailed structures in RVE, researchers have recently employed X-ray micro-tomography to directly obtain the geometry of the composites. This is a promising technique but is quite expensive and requires careful image processing. As an alternative, to achieve a fine balance between speed and accuracy in generating the RVE structure, we developed a novel 2-step geometrical modeling method in the invention with a one-time post-processing step to modify the local yarn geometry generated by TexGen.

As for the challenge to obtain unknown yarn properties at the mesoscale, a Bayesian model calibration and validation approach is developed for integrating the calibrated mesoscale stress emulator with macroscale part performance simulations. This is the first case in which Bayesian calibration and hierarchical multiscale techniques are utilized for simulation of uncured prepreg preforming.

The major technical target for the preforming modeling part of the invention is to develop a computational simulation method that can capture the deformation of carbon fiber composite prepregs, including part geometry, fiber orientation, and forming force during preforming, with high accuracy and less than 5% error. With this simulation method, the material cost and development period for design and optimization of carbon fiber composite prepreg preforming can be reduced significantly compared to the conventional manufacturing process design methods, which rely heavily on trial-and error experiments. This cost and time cut for the development of a carbon fiber composite manufacturing process will broaden the application of these advanced light-weight composites in the transportation industry, making a great contribution to the control of fossil fuel consumption and emission pollution.

For the success of development, experimental material characterization techniques are to be designed and performed systematically, first to provide correct input to simulation models. Then models at both mesoscale and macroscale are to be established and validated. The goal of mesoscopic modeling is to perform virtual material characterization to replace the unsatisfactory direct experimental characterization. The goal of macroscopic modeling, on the other hand, is to form a platform for part-scale preforming simulation when measured material properties are input. Finally, these modeling tools are to be combined with proper calibration techniques to form a high accuracy and high fidelity hierarchical multiscale modeling method for the prepreg preforming process.

Uncured Prepreg Characterization Experimental Methods

To obtain the well-defined constitutive model and the simulation schemes for the numerical calculation, characterization of the uncured prepreg is necessary. Several experimental methods to characterize the material have been developed in the invention: the uniaxial tension tests are used to characterize the tensile modulus along the fiber yarns; the bias-extension tests are used to characterize the shear modulus of the prepreg; the bending tests are employed to characterize the bending stiffness along the yarns; and the surface interaction tests are used to characterize the prepreg-prepreg and prepreg-tool interaction during the preforming process.

Uniaxial tension tests: Uniaxial tension tests are used to obtain the tensile modulus along the fiber yarns during the preforming process. The experiment setting is shown in FIG. 127 with the use of a tensile machine. The mechanical properties of the uncured epoxy in the prepreg is sensitive to temperature, so a temperature chamber was used here. Additionally, the digital image correlation (DIC) system was utilized to measure the strain distribution in samples.

In the uniaxial tension tests of the prepreg, the tensile modulus was determined from the fabric tensile test, not a yarn test, so that the tensile specimen could include as many unit cells as possible. Tests at the various temperatures that will be encountered during preforming were conducted. To avoid slippage between the specimen and the clamps caused by the viscous epoxy, the two ends of the specimen were cured before the tests to harden the material and ensure the clamping force during the tests.

Engineering stress and strain are used to normalize the load and displacement data. As an example, the curves at various temperatures are demonstrated in FIG. 128. It can be seen that as the temperature increases, the undulation stage of the material becomes longer, and the tensile modulus during the settle-down region slightly reduces. This is reasonable considering the softening of the epoxy at high temperature. Because of this phenomenon, the strain and the stress at the end of the undulation stage, and the stiffness after the undulation stage, were selected to properly describe the temperature effect on the uniaxial behavior of the prepreg.

Bias-extension tests: The bias-extension test determines the in-plane shear stiffness properties of the woven carbon fiber prepreg. The experiment setting is the same as that for the uniaxial tension test as shown in FIG. 127. To produce a pure shear central region, the fiber yarns were aligned ±45° off the loading direction, and the length of the specimen should be larger than twice the width in order to release the constraint along the yarn direction.

The temperature of the experiment was set at the various temperatures that will be encountered in preforming, which covers the range of the general preforming process temperature measured by the IR camera during the single dome preforming process performed in our lab. Various temperatures and tensile rates are included in the tests to investigate temperature and shear rate effects on the shear stress.

Some normalization methods are necessary for convenient utilization of the load-displacement data from the bias-extension tests and compensation for the size difference of the specimen. The engineering stress was used to normalize the load. As for the displacement, because the deformation is not uniform throughout the specimen, the normalization method was derived based on the assumption that the yarns are inextensible, and no slip occurs in the sample during the bias-extension tests. In order to validate this displacement normalization method, two bias-extension tests with different specimen sizes were performed. The parameters for the tests are listed in Table 9-1. The tensile rates are selected so that the normalized tensile rates are the same for both tests. From the results in FIG. 129, it can be seen that this displacement normalization method can properly compensate for the specimen size difference before shear locking when the large pure shear deformation happens in the central region.

TABLE 9-1 Parameters for the displacement normalization method in the bias-extension tests. Size Temper- Test (width*thickness*length) Tensile rate ature Short 50 mm * 1.1 mm * 104.31 mm 12.00 mm/min 60° C. specimen Long 51 mm * 1.1 mm * 136.00 mm 18.87 mm/min 60° C. specimen

As an example, the normalized load-displacement curves under various temperatures and shear rates are plotted in FIG. 130. It can be seen that the temperature plays an important role to the shear stress-strain relationship, especially at the temperature near 50° C., which is the point where the used epoxy begins to transfer from the solid state to the fluid state. The higher the temperature, the lower the measured stress. The possible reason for this is that at a higher temperature, the epoxy becomes softer, reducing the resistance for the fiber warp and weft yarns to rotate relatively. As for the deformation-rate effect, a higher deformation rate will lead to a higher final stress, which is reasonable because the deformation resistance caused by the viscosity of the epoxy would be larger in a higher deformation rate case. However, it should be noted that when the deformation is smaller, the increase of the load happens earlier at a smaller deformation. This phenomenon cannot be explained simply by the elastic or viscous behavior of the material. It may require further investigating in the future.

In order to further validate the kinematical assumption that the yarns are inextensible, and no slip occurs in the sample during the bias-extension tests, DIC technique was applied to examine the Green strain field in the real specimen, and the result is compared with the theoretical one derived based on the same assumption, as shown in FIG. 131. It can be seen that in the range where the DIC can work properly, the assumption holds well. When the shear deformation becomes large, however, the relative motion between the warp and weft yarns would scratch off the paint used for detecting the strain in the DIC system and the result can no longer be reliable. Advanced optical methods might be necessary in future work in order to validate the assumption further and get a more precise relation between the shear deformation and stress, which can help examine the strain field of the prepreg under large shear deformation during the bias-extension tests.

Bending tests: The bending stiffness of the material is also needed for proper simulation of the material behavior during the forming process. However, the softness of the prepreg under the preforming temperature makes it difficult to measure the bending stiffness via the standard 3-point bending test. As an alternative, bending tests in the invention utilize self-gravity of the specimen. In the test, one end of the rectangular prepreg sample was clamped horizontally on a support as a cantilever beam, as shown in panel (a) of FIG. 132. The prepreg would deform under gravity due to its low rigidity in the elevated temperature. The deflection of the sample tip was measured, and the deformed shape was analyzed by a digital image analysis. The entire system was placed in a temperature-controlled chamber, and the temperature was recorded. The prepreg deformation during the test in a 50° C. chamber temperature is given in panel (b) of FIG. 132 as a demonstration.

In this prepreg bending test, due to the strong geometric nonlinearity, the bending stiffness could not be directly calculated from the typical beam theory. As a solution, a simulation model was utilized to calculate the bending stiffness reversely. This simulation model utilized the homogeneous material properties. The compressive modulus of the material was modified until the same end tip displacement as the experimental result was reached. Then the effective compression stiffness could be obtained to properly describe the bending behavior of the prepreg.

Surface interaction tests: Before the invention, there was a lack of a systematic study for the characterization of the interaction between two different composite prepreg surfaces, even though this interaction surely affects the prepreg deformation and fiber orientation after preforming. To address this issue, a new experimental method was developed to characterize the interaction between uncured prepreg layers during the preforming process. The focus of this method is to characterize the influence of temperature, sliding speed, and fiber orientation on the tangential interaction. The apparatus, test procedure, and result of this new friction test method are to be illustrated specifically in the following part of this section.

The test apparatus developed in the invention is demonstrated in FIG. 133. The schematic of this apparatus is shown in FIG. 134. During the test, the top prepreg layer was clamped on the motion stage, which resembled the relative “pull-out” slip. The bottom prepreg was clamped on a stationary heating stage, which raised the temperature between the two prepreg surfaces. The edges of the top prepreg layer were fixed on the side of the motion stage instead of the bottom to avoid any possible edge effect, and to guarantee the surface-to-surface interaction during the entire test. A force-torque sensor was mounted on the motion stage to record both the normal force and the tangential force caused by sliding. In this test, the compression between the two layers was introduced by the displacement-control motion stage. Hence, the contact force varied during the test if the thickness of the prepreg across the entire test stage changed. Interestingly, as shown in the following analysis, this variation should not affect the interaction factor, which is calculated through normalizing the in-plane tangential force by the contact force.

An infrared (IR) camera, was also included in the system to measure the surface temperature distribution and provided the average values of a selected area. During each test, temperature was adjusted until it could be maintained within +1° C. variation from the desired value by carefully changing the power of the heating stage.

For this surface interaction test, the most important parameter is the surface temperature because it affects the viscosity of the resin in the composite. During the actual preforming process, the prepreg material was heated to 50 about 80° C. in a heating chamber and then placed in a press that is at room temperature. Thus, the temperature selected for the test ranges from room temperature (24° C.) to 80° C. The second considered parameter is the relative motion speed, because during preforming, 2D sheets are deformed into 3D parts. Then the relative motion speed between the material layers varies at different locations, resulting in a different interaction strength between the prepregs due to the shear rate effect caused by the resin viscosity at the interfaces. Finally, fiber orientation effect needs to be investigated. In industrial applications, prepreg layers with different fiber orientations are stacked together to optimize product performance in all directions. Since the surface texture of the composite is anisotropic, which affects the hydrodynamic interaction between the fabric and the resin, it is also important to test the material interaction subjected to different fiber orientation combinations.

The friction test results were analyzed when the tests with all parameter combinations were complete. The uncured prepregs are quite tacky in preforming because the resin is of high viscosity at the performing temperature. As a result, the Coulomb friction coefficient between two uncured prepreg layers can be higher than 1. To avoid confusion, the interaction factor similar to the Coulomb friction coefficient was defined to indicate the intensity of the yarn-resin-yarn interaction between the two prepreg layers. The results in FIG. 135 show that at the beginning of the test, large variations in the force components and the interaction factor value were observed. The reasons for this include: (1) the starting point of the sliding was at the edge of the bottom prepreg, which was not in the uniform temperature region because the heating stage only provided a uniform field at the center, and (2) during the first 30 seconds of the test, the two prepreg layers were in contact for heat conduction without relative motion. Thus, the prepregs were very tacky, resulting in a high initial interaction factor. Based on the steady temperature field measured with the IR camera image and the trend of the interaction factor curves, only the experimental data within the sTable 9-stage, which is the sliding distance ranging from 30 to 70 mm as indicated by the purple marks in FIG. 135, were utilized.

Periodical changes of the interaction factor are observed from the resulting plots, especially when the resin has very high viscosity. Take the interaction factor at the steady-state stage shown in FIG. 136, as an example. In addition to the small oscillations caused by stick-slip, the plots from all three tests show consistent and noticeable peaks and valleys that have similar amplitude and phase. This phenomenon will be further investigated by the hydro-lubricant interaction model in the next section.

The final results with respect to average stable stage temperature, relative motion speed, and fiber orientation combination are plotted in FIG. 137. Both the interaction and stick-slip strengths reach peaks values at 50° C., which is the critical temperature when the resin fully transforms from the solid to the fluid state and shows the highest viscosity. Below 50° C., the resin is in its solid state. It gradually becomes softer and tackier as the temperature increases, which leads to: (1) larger interaction factor because a stronger external force is needed to shear the resin in different layers under the relative sliding motion, and (2) larger amplitude and higher frequency stick-slip because of the more frequent molecular chain mixing and inter-diffusion, causing the tangential interaction force to fluctuate.

The temperature effect to prepreg surface interaction is that when the temperature is higher than 50° C., the resin fully transforms to a viscous liquid state and acts like a “lubricant” between the two prepreg layers. At this stage, further temperature increases reduce the resin viscosity and enhance its lubricity during sliding, resulting in a lower interaction factor, and less frequent stick-slip with a smaller amplitude. As for the relative motion speed effect, a weak influence at temperatures below 50° C. or above 60° C. was found. But the interaction factor has a positive relation to the motion speed at about 50° C., when the viscosity of the resin reaches the peak value. In regard to fiber orientation effect, when the temperature is higher than 50° C., the interaction factor becomes larger with the 0/90/0/90 fiber orientation, than that with the 0/90/−45/+45 orientation because the transverse fiber yarns in different layers are more likely to interlock with each other with the 0/90/0/90 orientation, making it more difficult for the layers to slide. This also explains the fact that the stick-slip strength is generally larger if the fibers in both top and bottom layers are aligned with each other, especially at 50° C. when the resin viscosity reaches the peak value. However, when the temperature is below 50° C. and the resin is in the solid state, the difference between the two orientation combinations becomes less significant. This is mainly due to the fact that the sheets are still in the solid state, and that the resin fills the surface “valleys” generated by the fiber yarns to flatten the prepreg. As a result, the orientation combination does not have a large effect compared to those at higher temperatures.

Hydro-Lubricant Uncured Prepreg Surface Interaction Model

The woven fibers in the invention form a certain texture of surface topography. It has a 2×2 twill structure, as shown both by the real material photo and the TexGen software model in FIG. 138. The characteristic sizes of woven structure, i.e., yarn width, yarn gap, and yarn thickness, listed in Table 9-2, were measured by microscopes from the cross-section of the material. The average thickness of the prepreg are obtained via a caliper.

TABLE 9-2 Parameters for the displacement normalization method in the bias-extension tests. Yarn width Yarn gap Yarn thickness Prepreg thickness 2.430 ± 0.112 mm 0.004 ± 0.004 mm 0.503 ± 0.012 mm 0.85 ± 0.15

Textures affect the interaction of textured surfaces. A hydrodynamic model was developed and applied to simulate and study prepreg surface interaction in the invention. In this model, the top and bottom woven fabrics were aligned to the same direction with 0/90/0/90 fiber orientation for 2D simplification. These fabrics were treated as rigid because 1) they were firmly stretched in the fiber matrix, so that the vertical deformation was minimal; and 2) the normal load was low. Relative movement of the interface can be considered by the general lubrication system illustrated in FIG. 139. This system is formed with two solids separated by a continuous fluid film. In the simulation, the solid geometry was determined based on the cross-section of the 2×2 twill prepreg, as shown in panel (b) of FIG. 135. The single warp yarn cross-section was idealized as an elliptical shape, while the cross-section of the weft yarn on top of the two warp yarns was modeled as a plane tangent to two half elliptical shapes. It was assumed that, in the simulation, the upper layer would move with respect to the lower one. To describe the dynamic of viscous resin, one-dimensional transient Reynolds equation for incompressible Newtonian fluid flow is utilized.

With this hydro-lubricant model, surface interaction at various temperatures is simulated at a relative motion speed set of 10 mm/s. The comparison of numerical and experimental results is shown in FIG. 140, where the average, maximum, and minimum values of the interaction factor are plotted. At 50° C., the numerically calculated interaction factor is significantly larger than the experimental one because the continuity assumption is not valid. In the simulation, the resin layer behaves like a continuous fluid with high viscosity, while in the experiment, the resin may only partially melt, so there is still an interface where friction takes place between the top and bottom prepregs. It should be noted that, at this interface which breaks the continuity assumption in the simulation, the friction should be lower. Moreover, in the numerical calculation, the prepreg fiber is assumed to be rigid for simplification. In the experiment, on the other hand, the highly viscous resin at this temperature leads to large fluid shear stress, deforming the prepreg fiber, changing the surface profile and in return reducing the interaction. At 60° C., the numerical results agree very well with the experimental ones because the viscosity falls in a reasonable range, and the continuity assumption is valid since the resin fully melts. At the 70° C. condition, the numerical predictions are slightly smaller than the experimental result. A larger interaction factor in the experiment is due to the direct contact between two woven fabrics. It was found that at this condition, the minimum film thickness would reach 0.06 mm during the calculation because the viscosity of the resin becomes very small. The minimum film thickness is in the same order of the 0.012 mm half-yarn thickness variation; thus, in the real tests, two woven fabric surfaces may contact each other at some positions, resulting in a boundary-mixed-hydrodynamic lubrication cycling.

For the interaction at 60° C., the numerical calculations with various relative motion speeds were then performed. The experimental and numerical results for the average, maximum, and minimum interaction factors are plotted in FIG. 141. The interaction model results agree well with the experimental ones in general. However, the speed effect is slightly more significant than that found in the experiments because of the hydrodynamics assumption between rigid surfaces in the model, which is sensitive to sliding speed. However, in the real experiment, other factors can also contribute to the speed effect. At low speed, there is sufficient time for the resin to mix and inter-diffuse, so that the resin is tackier and tends to stick the two surfaces together, thus increasing fluid resistance to motion. At high speed, the interaction force increases because of the viscous friction, so elastic deformation of the fiber increases correspondingly, which in return flattens the surface and reduces the interaction in the real materials.

Finally, with this hydro-lubricant model, the periodic interaction factor variation demonstrated in FIG. 136 was investigated. The Fast Fourier Transformation (FFT) was applied to both experimental and numerical results. Results at 60° C. and 10 mm/s are plotted in FIG. 142, showing that all the experimental and numerical curves have the 1st order length frequency of about 0.1/mm, which means that the interaction factor changes in the period of about 10 mm. This phenomenon is dominated by the size of the prepreg unit cell, which has a 2×2 twill element of 9.74 mm side length. However, for higher order frequencies, numerical results agree less with experimental ones, especially in terms of amplitude. This might be explained by the fact that viscoelasticity of the real material can absorb high frequency vibration energy during sliding.

Experiment validation demonstrates that under certain preforming conditions, i.e., 60° C. temperature and 5-15 mm/s sliding speed for the supplied woven prepreg, interaction between two prepreg surfaces can be explained by the hydro-lubricant mechanism and predicted via the numerical method developed in the invention. The elastic deformation of the fabric and the resin mixing with inter-diffusion at various deformation rates and temperatures should be considered in the future work in order to model the prepreg-prepreg interaction more accurately and predict the interaction behavior subjected to wider conditions.

Macroscopic Non-Orthogonal Material Model for Uncured Prepreg

In the invention a non-orthogonal material model for the CFRP preforming simulation was developed, aiming to accurately predict the deformation of the uncured prepreg during preforming especially under large shear. This material model was developed in the form of Abaqus® explicit user-defined material subroutine (Abaqus VUMAT) and LS-DYNA® user-define material subroutine (LS-DYNA UMAT). Because of its ease of use and high prediction accuracy for part shape and fiber orientation, this model was incorporated into the LS-DYNA® as MAT_293 (MAT_COMPRF). The fundamentals of this model can be found following this section.

Woven CFRPs have highly anisotropic mechanical properties, with large tensile modulus (10 GPa level) along warp and weft yarns because of the stiff carbon fibers reinforcement, but small intra-ply shear modulus (0.1 MPa level). During preforming, the most dominant deformation mode is the intra-ply shear. To capture this fiber-orientation-dominant anisotropy, the material model needs to simulate tension along the yarns and shear separately, even under large shear.

Stress analysis for the woven uncured prepreg with the non-orthogonal model developed in the invention is shown in FIGS. 143. σf1, and σf2 are the stress components caused by yarn stretch, and they are along the warp and weft yarn directions, respectively. σm1 and σm2 are the stress components caused by the yarn rotation. These stress components will be transformed into the local corotational coordinate, summed up as σXX, σXY, and σYY, and will then be output from the material model to FEM software. In this model, deformation gradient tensor F is utilized to trace yarn directions and stretch ratios during preforming via g=F·G, where g and G are the final and initial orientations of the local fibers respectively. It can be used to calculate α, which indicates the relative rotation between the local warp direction and the X-direction in the local corotational coordinate, and yarn angle β, which indicates the amount of shear deformation in the material.

This non-orthogonal material model was implemented into both Abaqus® and LS-DYNA®. This model enables users to directly input experimental data to define the stress-strain curves, as well as the shear locking angle, which indicates whether the shear deformation reaches to the extent that the rotation resistance between warp and weft yarns is no longer small compared to the tensile modulus of the material. FIG. 144 shows the calculation flowchart of this model in FEM software. From this flowchart it can be seen that warp and weft directions for each element are calculated from deformation gradient tensor. If the angle between warp and weft yarns are smaller than the shear locking angle, small shear modulus condition will hold, and total stress in the element will be updated via the non-orthogonal model. If the angle between warp and weft yarns reaches the shear locking one, resistance for further shear deformation will greatly increase because contacted fiber yarns stiffen the woven structure. In this situation, the “Yarn stretch caused stress” will still be calculated via the non-orthogonal model, while the shear components of the “Yarn rotation caused stress” will be updated with shear modulus of cured uni-directional (UD) carbon fiber composites.

Material characterization is essential for FEM models to accurately predict behavior of woven CFRPs during preforming process. It can be seen from FIG. 143 that the in-plane stresses caused by both yarn stretch and yarn rotation need to be calibrated for any specific woven material that is of interest. Their calibration can be performed directly by the uniaxial tension and bias-extension experiments. The out-of-plane behaviors of the uncured prepreg, are characterized by the single layer bending and double layer interaction tests.

When the material model and the experimental input are prepared, double-dome benchmark tests are conducted and simulated to validate the capability of the material model for a 3D shape forming, considering different yarn orientations and stacking sequences. These validation results indicate that this non-orthogonal model can partly reach the 5% error target for fiber orientation prediction.

Mesoscopic RVE Model for Uncured Prepreg

Mesoscopic RVE modeling and virtual material characterization with RVE requires building of an RVE finite element model, calibration of mesoscopic yarn properties, and generating a prepreg constitutive law as a function of strain. To build the mesh of a prepreg RVE with a fine balance between speed and accuracy, a novel 2-step geometrical modeling method was developed in the invention. In this method, the rough composite structure without yarn-to-yarn penetration is first generated by TexGen in Step 1 with the specified woven pattern and key characteristic sizes, such as weaving pattern, yarn width, yarn gap, and yarn thickness. Then, the mesh and the local yarn orientation corresponding to the structure is imported to a commercial finite element software in Step 2 to compress the structure in the thickness direction and satisfy the prepreg thickness requirements while maintaining the already assigned features. Finally, the deformed mesh and the local material orientation are exported to build the RVE for virtual material characterization.

This method was utilized to build the supplied 2×2 twill prepreg. Texture structure of the real 2×2 twill prepreg is shown in FIG. 138. The characteristic dimensions of the yarns and the average thickness data are listed in Table 9-2. Average values of yarn width, yarn gap, yarn thickness, and 2×2 twill pattern in TexGen were input first. Table 9-2 indicates that the yarn gap is very small compared to yarn thickness and width. To minimize yarn-to-yarn penetration, the shape of the yarn's cross-section was set to be lenticular. Thickness of the TexGen prepreg structure was also artificially enlarged from 0.85 mm to 1.2 mm to avoid penetration completely. The result is illustrated in FIG. 145 and as it can be observed, there is no longer any penetration between different yarns.

The drawback of this thickness enlargement, however, is that many gaps, as demonstrated in panel (b) of FIG. 145, are introduced in the structure. These artificial gaps significantly impair the prediction capability of the RVE model; upon exerting load, the inner gaps will greatly soften the RVE, reduce the response moduli, and elongate the undulation region. As a solution to close these gaps, the compression method is introduced in Step 2. To this end, two rigid plates are employed to compress the prepreg RVE in the thickness direction to reduce the thickness to 0.85 mm, which is the average value of the real material, as illustrated in FIG. 146. The rationality of this step is supported by the fact that there is no strict constraint for yarn cross-section shape and longitudinal path. At this stage, the mechanical properties of the prepreg yarn have not be characterized yet because they require calibration by the RVE, whose structure has not yet been obtained. As a result, in the compression simulation, the elastic moduli of the yarn are selected to be the same as the existing ones for cured unidirectional composites. The Poisson's ratios are set to be zero in all directions to avoid altering the yarn width due to the yarn deformation in the thickness direction. It should be noted that these yarn properties are only utilized to generate the RVE structure. They are not the same as the ones for prepreg obtained from Bayesian calibration.

In addition to the RVE structure, the yarn material model should also be correctly established. Because preforming is a one-step loading process where material recovery after the deformation is not neglected, yarns within RVE models are assumed purely elastic. Prepreg yarns that include quasi-unidirectional fibers and uncured resin exhibit a transverse isotropy. Direct implementation of such material behavior, however, leads to numerical errors. One kind of error happens when compression load is applied along the width direction to a single yarn. This loading condition is common for prepregs in shear deformation where, as illustrated in panel (a) of FIG. 147, fibers rearrange as resin flows in real yarns. Consequently, the yarn deforms (i.e., its dimensions change) while preserving the basic elliptical shape. In finite element simulation, yarns are treated as continuum with relatively flat cross-section geometry. If the transversely isotropic material model is utilized, numerical errors such as artificial bending and excessive element distortion will appear especially on the edges, as illustrated in panel (b) of FIG. 147. To address this issue, the transverse shear and normal behaviors in the yarn material model are decoupled to control bending and distortion of yarns while maintaining their compression property. With this approach, deformation similar to the real material can be achieved, as shown in panel (c) of FIG. 147.

Based on the decoupling approach, the yarn is modeled using an anisotropic elastic constitutive law with distinct Young's and shear moduli in different directions. This constitutive law is defined in the co-rotational frame which is updated with the deformation gradient tensor to accurately trace the local fiber orientation upon large yarn deformation and rotation under the RVE deformation. In the prepreg yarns, the very stiff carbon fibers are aligned in the longitudinal direction along which the applied load is predominantly present. Meanwhile, the soft uncured resin governs the transverse deformation. Therefore, it is straightforward to decouple the yarn deformation in the longitudinal and transverse directions.

Once the structure and the material model of the RVE are generated, they are input into the finite element simulation given normal true strain along yarns, shear angle, and yarn properties. After simulation, the stress of each element is extracted and averaged to obtain the stress response of the RVE. Mechanical properties of mesoscopic yarns including elastic moduli, Poisson's ratios, and friction coefficient are difficult to directly characterize because of small sizes, single yarn specimen preparation, and soft resin. As a result, the unknown material properties are manually adjusted at this stage and the stress prediction from the RVE is compared to the experimental data. One of the best example comparisons is illustrated in FIG. 148. The RVE result agrees very well with the experimental one when the shear angle is less than 0.6 radian, validating the 2-step approach developed. When the shear angle further increases, the discrepancy between the simulation and the experiment becomes large, indicating the necessities for calibration.

In the invention, Bayesian calibration is applied to obtain prepreg yarn properties for the first time. Uniaxial tension and bias-extension data is employed to: (1) estimate calibration parameters of the RVE model; (2) determine whether the RVE simulator is biased; and (3) build a cheap-to-evaluate emulator to replace the expensive RVE simulation in macroscale analyses. To this end, a modularized version of the Bayesian calibration framework of Kennedy and O'Hagan (KOH) is adopted. The goal of Bayesian calibration is to combine three data sources (experiments, simulations, and prior knowledge from experience in the field) to estimate the unknowns. As illustrated in FIG. 149, where x represents strain and θ represents yarn properties, this process starts by replacing expensive mesoscopic RVE simulation with a GP emulator (metamodel) η(x, θ) in Module 1. Then, uniaxial tension experimental data and prior knowledge on mesoscopic yarn properties p(θ) are used to fit the GP emulator δ(x) to the bias function in Module 2. Our reason for introducing δ(x) is that even if true calibration parameters were known (which they are not) and used in simulation, the stress predictions from the RVE model might not match with the experiments. In Module 3, joint posterior distribution of the mesoscopic yarn properties p(θ|d) are obtained given d, i.e., collection of the results from experiments and simulation. Finally, in Module 4, the updated emulator is compared against the bias-extension experimental data for validation. Once validated, the updated emulator, as virtually characterized constitutive law, is utilized to predict the stress response of the RVE under any strain state.

In practice, for supplied prepregs in the invention, uniform prior distributions (based on our experience) are chosen for θ which cover the entire range where η(x, θ) is fitted. Uniform prior distributions are preferred (over, e.g., normal distribution) since the range of the values that θ can take (and not, e.g., their most likely values) are known information. These ranges are chosen widely enough to ensure that the true (but unknown) calibration parameters are covered. Additionally, this choice guarantees that large variances are used to avoid diminishing the effect of the experimental data on the joint posterior distribution of θ.

TABLE 9-5 Prior and posterior distribution of the calibration parameters: The priors on θ = [E1, E2, μ] are uniform and denoted with Uni (lower bound, upper bound). Unlike the prior, the posterior distributions of the calibration parameters are neither uniform nor independent. Prior Distribution Posterior Mode E1 about Uni(20, 60)GPa, E2 about Uni(5,25)MPa, μ about Uni(0.15, 3) [ 46.8 GPa 23.5 MPa 1 . 3 ]

Calibrated yarn property results are shown in FIG. 150 and demonstrate that the marginal variances are relatively large, which was expected since (1) there are multiple sources of uncertainty such as experimental and simulation errors and simulator bias, and (2) limited data are employed in the Bayesian analysis: the calibration data is only 20 points from the uniaxial tension test while the RVE virtual testing goes into complicated loading conditions.

FIG. 151 illustrates the predictions of the orthogonal stress components by the updated emulator under various deformation states. Normal stress σ11 is plotted against normal true strain along warp and weft yarns, ε′11 and ε′22, for two different values of γ′12 in panel (a) of FIG. 151. Similarly, shear stress σ12 is plotted in panel (b) of FIG. 151 where its symmetry with respect to ε′11 and ε′22 is evident. Compared to σ12, σ11 is less sensitive to γ′12. It can also be observed that 012 monotonically increases as any of the strain components increase. This monotonic behavior is also observed in panel (a) of FIG. 151 but is slightly compromised when there is no shear strain (i.e., in the red surface). This small inconsistency may be due to (1) dynamic explicit numerical issues such as the artificially high strain rate to reduce the run-time in the RVE simulation, and (2) lack of simulation data with very small γ′12, resulting in extrapolation during the Bayesian calibration. In panel (c) of FIG. 151, uniaxial tension experimental and predicted results are plotted. Since this test was used for calibration, the predictions are expected to match the experiment. In panel (d) of FIG. 151 bias extension experimental and predicted results are plotted. Since this data are not used in calibration, FIG. 151 illustrates that the calibration has been effective in learning the stress-strain behavior. The posterior of the resulting GP model can now be used as the constitutive law of integration points in the macroscopic preforming simulations.

Multiscale Uncured Prepreg Preforming Model with Bayesian Calibration

As the constitutive law of the 2×2 twill prepreg with uncured thermoset resin, the mesoscopic stress emulator obtained from the virtual material characterization is implemented into the non-orthogonal material model to form a multiscale simulation approach. The flowchart of the developed approach is illustrated in FIG. 152. This emulator is learned at mesoscale and acts as the non-orthogonal material constitutive law by replacing the expensive mesoscale RVE finite element simulation at each integration point in macroscale preforming analysis. For the macroscopic constitutive law in this multiscale approach, deformation input includes normal true strain ε′11 and ε′22 along warp and weft yarn directions, and shear angle γ′12. These inputs are all calculated using the non-orthogonal coordinate algorithm. The predicted stress components are obtained in orthogonal material coordinate directly. Hence, the constitutive law does not require coordinate transformation of stress. It should be noted that the prepreg stress emulator is learned over the range of ε′11∈[−2, 2]%, ε′22∈[−2,2]%, and γ′12∈[0, 1] radian. For the deformation states outside these ranges, the prepreg will transfer into shear locking state, and the finite element simulation employs the shear locking state algorithm in the non-orthogonal model.

When the multiscale preforming simulation method is established, double-dome benchmark tests are conducted and modeled to validate the capability of the multiscale method for 3D shape forming considering different yarn orientations and stacking sequences. This validation result reveals that this multiscale method leads to a slight improvement regarding the prediction of part geometry and fiber angle distribution, with an average of 4.0% error for fiber orientation prediction, which achieves the proposal target. Moreover, the forming force prediction accuracy of this multiscale method sees a significant increase of over 26% compared to the experiment-based non-orthogonal model and it agrees very well with the experimental results.

FIG. 152 shows a flowchart of the developed multiscale preforming simulation method: The Bayesian calibration utilizes the RVE and experiments to obtain the yarn properties and the mesoscale stress emulator. The stress emulator is then implemented into the non-orthogonal material model for macroscopic preformation simulation.

The experimental methods to characterize mechanical properties of carbon fiber composite prepregs, the non-orthogonal material model for preforming simulation, the mesoscopic prepreg RVE finite element model with Bayesian calibration for virtual material characterization, and the multiscale simulation method for preforming developed in the invention leads to an accurate computational design and optimization approach for development of a CFRP parts preforming process. The high fidelity RVE simulation at the material structure level also provides insight guidance for the woven pattern and constituents design in composites. This approach is able to reduce the time period and material cost for the development of preforming compared to conventional trial-and-error methods, which rely heavily on real experiments. As a result, this approach enables researchers and engineers in both academic and industrial fields to invent and produce CFRP parts and corresponding manufacturing processes at a faster pace, at a lower price, and in a larger volume, broadening the application of these advanced composites and benefiting environmental emission and fossil fuel control.

The 2-step mesoscopic RVE modeling technique and multiscale simulation tool can also be commercialized and implemented into mature finite element software as a plug-in feature. For the RVE technique commercialization, the next necessary step is to integrate open TexGen software for the generation of fabric geometry/mesh with finite element software to form a complete package. As for commercialization of the multiscale tool, future needed steps include: (1) transferring the Bayesian calibration algorithm from current MATLAB code to Python or Fortran languages that can be utilized directly by finite element software; and (2) integrating RVE virtual testing model, Bayesian calibration, and part-scale preforming simulation model into a whole package.

In the preforming part of the invention, we invented a set of experimental and simulation methods that can speed up development of CFRP parts manufacturing at a low cost. This enables not only big companies, but also smaller research teams to design their own composites and corresponding manufacturing processes, which potentially increases applications and market needs for these advanced light-weight materials, reduces emission pollution from transportation industries, and motivates the carbon fiber composites production industry.

Carbon fiber composite prepregs are expensive and require special treatment both for storage (freezing) and for manufacturing (high temperature to melt resin, but not cure it). Systematically arranging physical experiments with proper deformation and temperature checking is very essential to avoid waste of raw material; and (2) Computation models, when hundreds of them need to be run for calibration, virtual material characterization, and design optimization, can be time consuming. Validation of single simulation cases in the aspect of mesh density, local material coordinate, material model, etc., is essential to ensure high efficiency for large scale computational simulations. As a summary, to save time, cut cost, and achieve neat and satisfactory results, either experiment or simulation research work requires sufficient preparation and planning, instead of simply performing them and relying on the trial-and-error method.

3. Mechanistic Continuum Models for CFRP

Carbon Fiber Reinforced Polymers (CFRP), including Unidirectional CFRP (UD CFRP) and twill woven CFRP (woven CFRP), have orientation dependent material properties. Simply put, the tension responses of UD and Woven CFRP will be different depending on the loading condition. This is due to the anisotropy nature of carbon fiber, which has different elastic moduli in fiber directions and in-plane directions. Hence, to study the mechanical behavior of CFRP material, it is a necessity to model CFRP's actual microstructure for analyzing the performance of cured CFRP, including UD and woven CFRP.

The multiscale method established in the invention provides tools to model CFRP microstructure by the RVE method. Numerical testing of RVEs predicts effective elastic material properties of UD and woven CFRPs. Based on a reduced order modeling approach on RVEs, a concurrent multiscale modeling method has been established to perform efficient part-level performance prediction.

Major achievements in the UD and woven modeling are summarized as below:

    • User-friendly UD RVE modeling package.
    • User-friendly woven CFRP modeling package.
    • Prediction of UD elastic constants with less than 10% difference to test data.
    • Prediction of woven elastic constants with less than 10% difference to test data.
    • Uncertainty quantification (UQ) for Woven RVE.
    • Novel multiscale modeling concurrent with modeling for UD CFRP.
    • UD part performance predictions with less than 10% difference to test data.
    • Woven concurrent modeling.

The properties of CFRP composites is anisotropic and microstructure dependent. Therefore, accurate capture of CFRP's mechanical properties, such as stiffness tensor, requires 1) reconstruction of the microstructure; and 2) accurate numerical modeling. For UD CFRP, its stiffness tensor can be simplified as the volumetric average by the rule of mixture. It is possible to estimate the UD stiffness tensor by using either Voigt average or Reuss average, but the accuracy is questionable since Voigt average gives the upper limit while Reuss average gives the lower limit. In addition, it is also proposed the modeling of UD CFRP by assuming a well-structured and periodic packing pattern of fibers, such as hexagonal packing, and model UD CFRP by Representative Unit Cells (RUC). RUC provides easy modeling of UD CFRP since it only models the minimum repeating unit in UD CFRP and it allows different packing patterns and fiber volume fractions. RUC can be easily modeled in finite element mesh and allows one to compute effective UD elastic properties without dealing with algebra, compared to the analytical homogenization approach. Unfortunately, in real UD CFRP product, carbon fibers are of a random distribution. Therefore, the RUC model might not provide accurate information about the UD's properties. For woven CFRP, RUC can be used for modeling mechanical properties of woven composites. However, due to the aforementioned assumption, RUC is not the ideal way of modeling woven CFRP when one wishes to capture certain microstructure variation. In order to include realistic microstructure, a better modeling technique needs to be identified and implemented so the properties of UD and woven CFRP can be predicted.

In the invention, the Representative Volume Element (RVE) approach is adopted to faithfully model UD's microstructure. The RVE model of UD captures the random distribution of fibers in the matrix material and can be used for finite element analysis. Building a UD RVE allows one to consider arbitrary fiber distribution and fiber shapes. It is expected that UD RVE can give a good prediction of UD's mechanical properties compared to test data. For woven composites, the RVE model allows one to capture a larger region with multiple fiber tows in warp and weft directions, and an accurate prediction of woven mechanical performance can be made. In addition, woven RVE enables uncertainty quantification of woven CFRP, where the effect of fiber volume fraction and fiber misalignment in fiber tow can be quantitatively analyzed. Hence, the RVE approach provides a clear structure to the property map between the UD and woven composites and their mechanical properties. The current scope is to use RVEs for accurate prediction of elastic stiffness tensors for both UD and woven composites.

Moreover, the UD RVE can be applied into a recently proposed Reduced Order Modeling (ROM) method, namely Self-consistent Clustering Analysis (SCA). SCA allows one to compress UD RVE from many voxel elements into a ROM database made with several clusters. The ROM of the UD can be solved using the SCA method, hereafter called “UDSCA”, to compute elasto-plastic responses of UD in an efficient manner. UDSCA not only provides an efficient way to compute mechanical responses (elastic and plastic) of UD, but also links UD microstructure to UD part performance. A concurrent multiscale modeling framework is established for UD material for the first time and it can be used for structural property prediction.

The present goal of UD and Woven modeling is to develop a complete modeling workflow that allows the user to generate UD and woven microstructures and extract elastic material properties, especially the stiffness tensors. It allows one to build a UD or Woven microstructure as an RVE in a finite element mesh. Microstructure information, such as UD fiber volume fraction and yarn orientation in woven composites, can be assigned by the user. The finite element mesh can be used to perform traction free loadings in three normal directions and three shear directions. RVE effective stress and strain results from all six loadings will be used to compute stiffness tensor, in a 6 by 6 matrix. The present process provides direct numerical homogenization of CFRP's material properties. Young's moduli and shear moduli can be computed and compared against experimental results. The expected different between UD and Woven elastic constants from RVE models and test data is less than 10%.

Moreover, an uncertainty quantification workflow is established, for the first time, for woven CFRP. In this workflow, variations of the woven microstructure can be modeled in woven CFRP. Several microstructure variations, such as yarn angle, fiber misalignment, and fiber volume fraction are considered. The effect of those variations on effective woven elastic material properties can be measured quantitatively.

For UD and Woven CFRP, an efficient reduced order modeling approach, namely self-consistent clustering analysis (SCA), is applied to reduce the computational cost of RVE computation. This allows one to compute the responses RVEs on-the-fly and enables a concurrent multiscale modeling framework. The multiscale modeling framework establishes a concurrent multiscale modeling framework where prediction of macroscale structure performance is made possible. Test cases of UD CFRP structure will be presented with experimental validation. It has broad potential in the evaluation of CFRP structure performance through numerical models and can be used for future CFRP structure design.

UD RVE Modeling

In the ICME process, a bottom-up by a multiscale modeling approach is adopted for CFRP. As shown in FIG. 153, it is convenient to use a three-scale model to describe a cured CFRP part: UD in microscale, woven in mesoscale, and part in macroscale. The RVE models for UD and woven are built in order to evaluate their mechanical properties, such as stiffness tensors. The information obtained at lower scales can then be used in higher scales. For example, fiber tow in the woven RVE can be treated as having the same properties of UD RVE. Therefore, the UD RVE can also be used to compute fiber tow properties that are used in evaluating woven RVE's mechanical properties. Therefore, ud modeling should be introduced first.

An image of cured UD microstructure used in the invention is given in FIG. 154, where random fiber distribution is observed. The fiber volume metric fraction of all UDs in the project is 51%. Hence, the modeling of UD requires one to build a numerical model that captures random fiber distribution with a given fiber volume fraction. For that purpose, UD RVE is used to model microstructure observed in FIG. 155. Key assumptions for UD RVE modeling are given as below:

    • 1) 51% overall fiber volume fraction;
    • 2) Fiber has a circular shape with a diameter of 7 μm;
    • 3) Fibers are perfectly straight;
    • 4) RVE has a square cross-section, with a length larger than 70 μm.

The process of building a UD RVE is similar to packing multiple fibers into a square domain. Since all fibers are assumed to be perfectly straight, it is convenient to reduce the 3D domain into a 2D domain, where multiple circles are packed into a square. This 2D domain can be extruded in the fiber direction to form the final 3D UD RVE. Here, the UD RVE geometry is 84 μm by 84 μm by 2.8 μm, as shown in FIG. 136. In FIG. 156, fiber direction is defined as 1 direction and two transverse directions are defined as 2 and 3, following the right-hand rule. A UD RVE package is made for this particular purpose so the user can build new RVEs with given fiber diameter and arbitrary RVE size. The UD RVE shown in FIG. 136 is discretized by cubic voxel element with 600 elements in 2 and 3 directions and 20 elements in 1 direction. When the UD RVE is used for finite element analysis, each voxel element contains only one integration point.

The UD RVE can be used to predict elastic stiffness tensor of UD. The material properties of fiber and matrix are given in Table 9-6. Since the primary focus is computing elastic material properties of the UD, only elastic material properties of fiber and matrix are needed.

TABLE 9-6 Fiber and matrix properties of cured woven composites Fiber E11 E22 E33 G12 G13 G23 ν12 ν13 ν23 245 GPa 19.8 GPa 19.8 GPa 29.2 GPa 29.2 GPa 5.92 GPa 0.28 0.28 0.32 Matrix E ν 3.79 GPa 0.39

To compute the UD stiffness matrix, it is necessary to review the definition of general monoclinic material. The general strain and stress relationship of monoclinic material is given as below, where Voigt notation is used. For UD RVE, due to its anisotropic nature, we have: S12=S21, S13=S31, S23=S32, S22=S33, S44=S55, S14=S41=0, S24=S42=0, S34=S43=0, and S56=S65=0. To compute individual entries of the S compliance matrix, one needs to perform loading on the UD RVE such that only the stress in the loading direction is non-zero. This so-called orthogonal loading condition allows one to compute S column by column. Therefore, loadings in 11, 22, 33, 12, 13, and 23 directions need to be performed. Once the full compliance matrix is constructed, the stiffness tensor is merely the inverse of the compliance matrix.

[ ɛ 1 1 ɛ 2 2 ɛ 3 3 γ 1 2 γ 1 3 γ 2 3 ] = [ S 11 S 1 2 S 1 3 S 1 4 0 0 S 2 2 S 2 3 S 2 4 0 0 S 3 3 S 3 4 0 0 S 4 4 0 0 symmetry S 5 5 S 5 6 S 6 6 ] [ σ 1 1 σ 2 2 σ 3 3 σ 1 2 σ 1 3 σ 2 3 ] = [ S ] 1 - 2 - 3 [ σ 1 1 σ 2 2 σ 3 3 σ 1 2 σ 1 3 σ 2 3 ] (9-8)

Once the UD RVE is generated, the mesh can be used in FE software, such as ABAQUS, to perform loadings in six loading directions mentioned above. Since both fiber and matrix are assumed to be elastic materials, the UD RVE responses will be strictly elastic. The UD effective stress and strain are computed using Eq. (9-9) shown below, where σmicro and εmicro are stress and strain tensors of each voxel element in the UD RVE.

σ ¯ = 1 | V | V σ m i c r o d V , ɛ ¯ = 1 | V | V ɛ m i c r o d V (9-9)

Once σ |ε are computed, one column of the compliance matrix can be computed according to the specific loading direction. For example, in the 11 loading direction, all stress components will be zero except for σ11. By performing basic matrix algebra, the first column of the can be computed entry by entry.

The quantities of interest of the UD RVE are the Young's and shear moduli, as well as Poisson's ratios in different directions. The conversion between UD compliance matrix and elastic moduli can be found using the following equivalence:

{ S 1 1 S 1 2 S 13 S 2 1 S 2 2 S 2 3 S 3 1 S 3 2 S 3 3 S 44 S 55 S 66 } = { 1 E 11 - v 21 E 22 - v 31 E 33 - v 12 E 11 1 E 22 - v 32 E 33 - v 13 E 11 - v 23 E 11 1 E 33 1 G 12 1 G 13 1 G 23 } (9-10)

The elastic material constants of UD RVE is summarized in Table 9-7, along with the experimental results. Most of the differences are all within 5% of the experimental data, outperforming the original proposed target of this work. Note that there has a relative large difference for shear moduli G12 and G13 between the experiment and prediction, but it is still within the target of 10% difference.

TABLE 9-7 Comparison of UD elastic moduli computed by UD RVE and experimental data. E11 E22 E33 G12 G13 G23 v12 v13 v23 UD RVE 127.92 GPa 8.56 GPa 8.60 GPa 4.28 GPa 4.35 GPa 2.66 GPa 0.33 0.33 0.60 Experimental 133.87 GPa 8.89 GPa 8.89 GPa 4.66 GPa 4.87 GPa 2.62 GPa 0.32 0.32 N/A Data Difference, % 4.44 3.80 3.25 8.17 10.68 1.45 3.16 2.75

The current package assumed fiber geometry in a circular shape, but it can be of an arbitrary shape. Also, the present circle packing method is considering complete random fiber distribution, but this limits the maximum fiber volume fraction to be 60%. To achieve higher fiber volume fraction, the algorithm needs an extra function that can rearrange fiber locations in order to exceed the 60% limit.

The generated UD RVE mesh usually has a considerable number of voxel elements, more than 1e6. Hence, UD RVE computation is most suitable for High Performance Computing cluster when finite element method is used. Since the UD RVE is discretized by voxel elements, this voxel mesh is essentially a 3D image. The Fast Fourier Transformation (FFT) based homogenization scheme is a favored algorithm to input voxel mesh and compute the overall stress and strain responses. If the FFT based homogenization method is adopted, one can compute RVE elastic responses using a single workstation.

In short, the UD RVE model has established a convenient work-flow that allows one to build UD RVE with desired fiber volume fraction. All predicted UD elastic properties either met or exceed the project requirement of 10% in difference.

Woven RVE

The woven RVE generation utilized TexGen, an open source software that allows one to build a textile structure at any given pattern and fiber tow (or yarn) geometry. In the invention, the woven CFRP is made of twill woven. The minimum repeating unit of the twill woven includes four wrap and four weft fiber yarns. The woven RVE generated is shown in FIG. 156, where the woven RVE is discretized by voxel elements with a resolution of 210 by 210 by 20. Wrap yarn is in the 2 direction and weft yarn is in the 1 direction.

Once the mesh of woven RVE is generated, it can be used in FE software to perform numerical homogenization to obtain its elastic material constants. Here, the matrix material has the same material properties. Fiber yarn property is assumed to be the same as UD CFRP with 65% fiber volume fraction. Due to the aforementioned limitation, the fiber yarn properties are computed using the analytical approach.

To analyze the elastic responses of the woven RVE, the matrix forms the stiffness tensor (due to the usage of Voigt notation) needs to be computed using six orthogonal loading conditions, same as the UD RVE. The stiffness matrix can be computed conveniently once the compliance matrix is computed. The only difference is that for woven RVE all components of the compliance matrix listed in Eq. (9-8) need to be computed.

The computed effective elastic properties of woven RVE are listed in Table 9-8 below. All prediction met with the proposed 10% difference compared with experimental data.

TABLE 9-8 Woven elastic moduli from woven RVE and experimental data. E11 G12 G23 Woven RVE 59.96 GPa 5.68 GPa  3.6 GPa Experimental Data 65.95 GPa 5.18 GPa 3.49 GPa Difference, % 9.08 9.65 3.15

The advantage of using a woven RVE numerical model is that various microstructure uncertainties can be addressed in the RVE model and quantitative analyses can be done to understand the effect of those uncertainties. In this work, uncertainty quantification for woven CFRP is introduced for the first time. The uncertainty quantification allows one to address uncertainties resulted from various manufacturing processes, such as pre-forming and curing. By giving a quantitative measurement of uncertainty effect, it is possible to link manufacturing process to the final CFRP performance, which is an important part of the ICME process. Here, the woven RVE also be used to examine the effect of three woven microstructures as shown in FIG. 157: 1) yarn angle; 2) yarn fiber volume fraction; 3) yarn local fiber misalignment.

The effect of yarn angle is studied by constructing woven RVEs with various yarn angle α shown in FIG. 157. The general assumption in the past woven CFRP research is that woven will maintain a 90° yarn angle, or the orthogonal configuration, after the performing and curing processes. However, it is shown that yarn angle would vary across the woven CFRP part after the curing process. Therefore, non-orthogonal woven RVE where yarn angle is less than 90° needs to be studied. Here, the stiffness matrices of each woven RVE realization are computed and all components are plotted in FIG. 156. Note the local material orientation is the one shown in FIG. 156.

From FIG. 158, the yarn angle has the most significant effects on C11, C22, C44, and C24. As the yarn angle decreases, wrap yarns will gradually leaning to 1 direction and C22 direction reduces significantly. However, C11 remains constant until the yarn angle is less than 60°. This means that the interaction between wrap and weft yarns is not significant when the yarn angle is larger than 60°. C44 tends to increase as yarn angle decreases, this means the woven in-plane stiffness at a smaller yarn angle will be stronger than that at a larger yarn angle. The concave shape of C24 reveals the shear-tension coupling effect, where the in-plane shear strain will contribute to stress in 22 direction when yarn angle is not 90°. All those observations reaffirm the importance of woven yarn angle for an accurate capture of the mechanical properties of woven RVE.

Besides yarn angle, yarn fiber volume fraction, denoted as Vf, and yarn fiber misalignment effects are also investigated. The results obtained in Table 9-8 consider neither uncertainty, meaning the yarn material is homogenous, which is usually not the case for real material manufactured due to manufacturing process variations. As shown in FIG. 156, each yarn in the woven RVE is made with multiple voxel elements, where each voxel element contains one integration point, representing one UD RVE. Therefore, by varying fiber volume fraction at each voxel element, inhomogeneity can be added to the yarn. By assuming fiber volume fraction on each yarn is following a Gaussian distribution, it is possible to assign different UD properties in different voxel elements to simulation different Vf in the yarn. The effect of Vf is summarized in Table 9-9, where Vf is following a Gaussian distribution with mean Vf=65% and variance σvf2=0.09%. It can be concluded that the effect of σvf2=0.09% does not affect woven property significantly, primarily depending on its mean value.

TABLE 9-9 Effect of yarn fiber volume fraction on homogenized material properties. E11 (GPa) E22 (GPa) E33(GPa) G12(GPa) G13(GPa) G23(GPa) Vf = 65% 59.96 59.96 12.57 5.68 3.60 3.60 Vf = 65%, 59.96 59.94 12.60 5.66 3.58 3.58 σVf2 = 0.09% Percentage 0.0011 0.0218 0.2128 0.4066 0.4436 0.4425 Difference (%)

Moreover, each voxel element contains a local material orientation that aligns with the yarn center line for homogeneous material. Fiber misalignment is considered as the deviation from perfect alignment direction. Shown in FIG. 157, vector {right arrow over (g)}1 represents the direction of perfect fiber direction, which is essentially the tangent line of yarn center line. Plane {right arrow over (g)}2{right arrow over (G)}3 is the yarn cross-section and {right arrow over (g)}1 is orthogonal to the plane.

Angle θ (0°≤θ≤90°) and Φ (−180°≤Φ≤180°) are used to establish misaligned fiber direction {right arrow over (f)}1, {right arrow over (f)}1, {right arrow over (f)}2, and {right arrow over (f)}3 represent transverse isotropic material frame accounting for fiber misalignment. Equations for calculating {right arrow over (f)}1, {right arrow over (f)}2, and {right arrow over (f)}3 given as below:

f 1 = g 2 g 2 sin θ cos Φ + g 3 g 3 sin θsinΦ + g ¯ 1 g 1 cos θ f 2 = g 2 g 2 cos θ cos Φ + g 3 g 3 cos θ sin Φ - g 1 g 1 sin θ f 3 = f 1 × f 2 (9-11)

For fiber misalignment, θ and φ follow gaussian distribution by letting mean θ=10°, variance σθ2=2, mean φ=0°, and variance σφ2=2500°2 (to make sure for all element within the yarn, its φ will fall between −180° and 180° following three sigma rule).

In summary, woven RVE modeling provides a straightforward numerical analysis tool for studying woven CFRP mechanical properties, where all predictions are within 10% difference compared to test data. Moreover, the test cases of woven uncertainty parameters illustrate the microstructural effect in woven CFRP. Woven UQ provides a convenient numerical solution to evaluate possible uncertainties caused by different manufacturing processes.

Reduced Order Modeling of UD CFRP

Aforementioned ud package is able to generate a UD RVE in voxel mesh and allow the user to analyze the mechanical responses of UD CFRP. However, the high computational cost associated with the fine voxel mesh requires certain reduced order model (ROM) techniques to achieve 1) faster RVE responses computation; 2) linking UD RVE to large-scale part-level model for part performance prediction with experimental validation (within 10% difference).

TABLE 9-10 Effect of Fiber Misalignment on Homogenized material properties. E11 (GPa) E22(GPa) E33 (GPa) G12 (GPa) G13 (GPa) G23 (GPa) No- 59.96 59.96 12.57 5.68 3.60 3.60 Misalignment θ = 10°, 48.01 47.14 12.42 6.15 3.67 3.68 σθ2 = 2 Φ = 0°, σΦ2 = 2500 Percentage 24.89 27.19 1.17 7.69 1.88 2.15 Difference (%)

A recently proposed reduced order modeling method, namely self-consistent clustering analysis (SCA), is a promising method for building ROM for arbitrary voxel mesh, including UD RVE. It is based on the FFT homogenization scheme. In FFT homogenization scheme, strain tensor at each voxel is treated as a combination of overall imposed strain εMacro and a polarization strain {tilde over (ε)}, shown in Eq. (9-14) below


ε(X)={tilde over (ε)}+εMacro  (9-14)

Above equation, also known as Lipmman-Schwinger equation, allows one to solve local strain responses ε(X) when εMacro is fixed. This is the basic of fft homogenization method, which is time consuming since the evaluation happens for all voxel elements. Eq. (9-3) can also be written as Eq. (9-15) as below


εMacro−ε(X)−∫ΩΓ0(X,X′):[σ(X′)−C0:ε(X′)]dX′=0,X∈Ω  (9-15)

Liu et. al. proposed a reduced order modeling approach by re-discretizing the voxel mesh into several clusters. Assuming the original voxel mesh contains N elements, the mesh can be compressed into K clusters, where N>>K. Eq. (15) is reformulated as Eq. (9-16) as shown below.

ɛ M a c r o - ɛ I - J = 1 K D IJ : [ σ J - C 0 : ɛ J ] = 0 (9-16)

Eq. (9-16) can be easily solved using Newton's method. Since N>>K, Eq. (9-5) is a much smaller linear system to solve than Eq. (9-15).

To apply SCA to UD RVE, the first step is to build the UD RVE database. This involves two steps:

    • 1. Compressed original RVE from voxel mesh into clusters.
    • 2. Compute interaction tensor DIJ between all cluster pairs.

Once the RVE is compressed, each voxel will be labeled with a cluster. This is illustrated in FIG. 159 where the RVE is decomposed into 10 clusters: 2 in the fiber phase and 8 in the matrix phase.

Once the UD database is built, Eq. (9-16) is solved to compute stress and strain responses in each cluster when an external loading is given. This rom, hereinafter referred as udsca, can be used to compute elasto-plastic responses of UD RVE in a timely fashion. A numerical verification of UDSCA is performed as shown in FIG. 55, where transverse tensile loading is considered. In this verification case, two different ROM resolutions are used: one with 16 clusters in the matrix phase and the other with 8 clusters in the matrix phase. The number of clusters in the fiber phase is kept as two. The result showed that using 8 clusters in the matrix phase and 2 clusters in the fiber phase provides good accuracy comparing to DNS solution. Hence, this ROM is used for all UD concurrent modeling cases.

For UD 2-scale concurrent modeling, it follows the schematic shown in FIG. 160. The macroscale part is discretized as finite element mesh. The ROM of UD RVE intakes strain at the integration point and then passes back stress response to the integration point.

UD Off-Axial Coupon Tensile Concurrent Modeling

Next, UDSCA is applied to a coupon off-axial test model to perform concurrent multiscale modeling. For a realistic representation of the epoxy matrix, a paraboloid yielding surface is applied, where the tension and compression curves are extracted from FIG. 45.

Through the coupon test cases, two important problems are addressed:

    • (1) Computing material microstructure evolution on-the-fly by realistic RVE.
    • (2) Prediction of CFRP part performance using the multiscale method.

For the coupon model, exact geometry from NIST is used, as shown in FIG. 161. The coupon model is built in commercial finite element software LS-DYNA. Note the teal region is the UD laminate made with 12 UD laminae. It is impossible to model every single fiber in the coupon explicitly since at least 409,422 carbon fibers need to be modeled. If a finite element mesh shown before is coupled to individual integration point of each finite element in the coupon mode, the computational cost is still huge, and the estimated solution time is beyond the capability of existing HPC. However, using UDSCA, UD RVE responses at each integration point can be computed in an efficient manner.

The simulation took 2560 CPUS hrs to complete. The stress vs. strain curve in the y-direction is computed and compared with the experimental result, as shown in FIG. 1253. A summary of the coupon test is given in Table 9-11, where the concurrent model is able to predict ultimate stress and strain reasonably well.

FIG. 9-15. Normal Stress and Strain Curve of UD off-axial Coupon Test.

TABLE 9-11 Comparison of Ultimate Normal Stress and Normal Strain. Prediction Experiment Difference Ultimate Normal 404.809 MPa 395.639 MPa 2.3% Stress Ultimate Normal 0.011 0.0118 6.7% Strain

In addition, FIG. 162 shows the von Mises stress of local RVEs and the coupon before crack initiation and after crack formation. In FIG. 162, RVEs that represent four different integration points are visualized. In RVEs representing integration points on the crack, stress becomes zero as the integration point is deleted from the coupon model. In RVEs representing integration points that are not deleted, stress is still nonzero due to stress wave propagation. The concurrent capture of macroscale and microscale stress evolution is made possible by the concurrent multiscale modeling scheme.

UD Crash Test Concurrent Modeling

The UD crash test setup is shown in panel (a) of FIG. 163, where the model is a quarter model based on the model provided by ford. The same concurrent scheme shown in ud coupon test is used.

The impactor force vs. Time is shown in panel (b) of FIG. 163 up to 4 s. It can be concluded that the time interval between the 1st peak force and the 2nd peak force predicted by the concurrent model is similar to that of the test data. The peak force recorded is given in Table 9-12, with a difference of 13.2%.

TABLE 9-12 Comparison of Peak Force Peak Force Prediction   11 × 1e4 kN Test 12.68 × 1e4 kN

UD Dynamic 3-Pt Bending Concurrent Modeling

The udsca is also applied to the ud hat-section 3-pt bending model. The model is shown in FIG. 164, along with the schematic of information exchange between the model and lower level UD RVE.

After a total displacement of s, failure in UD laminae is observed. Peak load and peak impactor acceleration are reported in Table 9-13, where the difference is within 10% compared to test data.

TABLE 9-13 Comparison of Peak Load and Peak Impact Acceleration Peak impactor Peak load(n) acceleration (m/s2) Experimental data 10328.1 0.390 Concurrent simulation 9660.0 0.382 Difference, % 6.47% 2.05%

Reduced Order Modeling of Woven CFRP

A 3-scale concurrent modeling for woven RVE has been established. The scheme is illustrated in FIG. 166 below. In FIG. 165, stress response at the macroscale integration point is computed by the ROM of woven RVE using sca method. For each yarn cluster of woven RVE, a ROM of UD RVE is assigned and solved by SCA. Considering the information across three different scales, this scheme is called 3-scale concurrent modeling. Note that there is no validation plan for 3-scale concurrent modeling in the invention. Numerical samples are shown to illustrate the concept.

The geometry details of woven RVE is shown in FIG. 166 where all parameters are provided by cao group. In the woven RVE, the volume fraction of yarn is 77% and volume fraction of fiber in each yarn is 60%.

A comparison between 2-scale and 3-scale single element simple shear tests is performed. Stress and strain results are shown in FIG. 167. It is reasonable to conclude that 2-scale model predicts stiffer shear responses of woven RVE since yarn phase is considered as elastic materials. The 3-scale model will require more computation cost due to the consideration of UD RVE and it can be used when woven exhibits highly nonlinear responses due to plasticity.

If the yarn nonlinearity (such as plastic behavior) is not of interest, it is also possible to replace UD RVE with a set of elastic constants of UD with a fiber volume fraction of 60%. This will reduce 3-scale concurrent model to 2-scale concurrent model for woven, where the only matrix is modeled as elasto-plastic material. An orthogonal woven biaxial tension test shows almost linear stress and strain curves. Hence, a 2-scale concurrent modeling of woven bias tension simulation is performed. The test setup is shown in FIG. 168. The same geometry is used in the concurrent modeling of the woven bias sample. Note here the woven sample is made of single layer of woven microstructure shown in FIG. 166. The final σ22 contour and σ22 vs. ε22 plots are given in panels (a)-(b) of FIG. 169, respectively. In panel (b) of FIG. 169, the σ22 vs. ε22 are the averaged values obtained from those elements in the gauge zone shown in FIG. 168. The predicted τ12 is slightly lower than test data, but it shows the same trend. The difference might be due to various factors, such as inaccurate woven geometry and yarn properties. Further investigation is required.

Benefits Assessment

The UD and woven CFRP RVE modeling packages provide alternative solutions to investigate CFRP mechanical properties. It can be applied to different constituents and predict elastic stiffness tensors of UD and woven composite. The UQ function of woven RVE allows one to link manufacturing process parameters to final product performance.

Moreover, the UD and woven concurrent multiscale modeling provides an accurate and efficient prediction of part-level product performance. It can be applied to a virtual verification platform where the concept design is evaluated. It allows optimization of the new design and can significantly reduce the number of costly experiments. The potential clients of this technology can be broad, as long as there is a need for developing new composite applications.

In the multiscale modeling work, it is understood that microstructure plays an important role in modeling CFRP materials. An efficient reduced order modeling method sca is introduced to integrate CFRP microstructure (UD and woven) into the part-level model to predict structural performance.

The next step will involve investigation of modeling of UD dynamic problems, such as ud hat-section crash and dynamic 3-pt bending. Current models suffer from numerical instability due to high loading rate. It is expected to use different stabilization methods to improve the concurrent scheme for better accuracy.

4. Stochastic Multi-Scale Characterization

Our research enables investigating the variability of part properties and behavior as a function of uncertainty sources at multiple length-scale and, subsequently, identifying the most important uncertainty sources that should be monitored during manufacturing. The developed methods and tools enable modeling spatiotemporally with varying uncertainty sources and, additionally, couple structural and material-related uncertainties across different length-scales. We achieve these by introducing the Top-down sampling method that builds nested multi-response Gaussian processes to parsimoniously quantify the random fields and, hence, the underlying physical uncertainty sources. Our approaches can be easily used to conduct sensitivity analyses for dimensionality reduction, i.e., identifying the most important uncertainty sources as well.

Compared to prior model-based UQ research, the UQ study of UD composites in the invention is image-based and microstructure-oriented. Two sources of uncertainty are considered: fiber waviness and fiber spatial distributions, both of which can be characterized from microscopic images provided by Ford. Machine learning and applied statistics methods are utilized to develop image analysis tools to extract information about the variations of the uncertainty sources, and generative statistical models are constructed for generating realistic random samples with variations mimicking our observations from the image data. For fiber spatial distribution, tree regression is used for image characterization and a hierarchical nonparametric sampling method is developed to sample the realizations from a nonstationary and nonhomogeneous RF. The local fiber waviness is obtained from images through a specially designed segmented regression algorithm, and new random samples are generated via a frequency domain time series analysis approach. Finally, the joint sampling of the two quantities, in which spatial constraints exist, are discussed and the corresponding codes are implemented.

The developed computational method and tools are applicable to many material systems and the corresponding multiscale material simulators. We have demonstrated their effectiveness in modeling uncertainty sources in unidirectional and cured woven composites. Our computational methods and tools can be validated against experimental results once they are available. In the case with cured woven composites, we demonstrate how various uncertainty sources such as yarn angle and fiber misalignment, which are introduced at, respectively, mesoscale and microscale, can affect the part performance during operational conditions. Our results indicate that, even in linear analyses, such uncertainty sources could have significant impacts on the results. With the UD UQ tools, random samples that represent the variations in the real UD material can be generated for further computational study, including their impact on material properties and part performance.

Our contributions are the first to investigate the uncertainty in multiscale material simulations which allows the systematic study of the effect of uncertainty to (i) engineer more reliable materials, and (ii) reduce the manufacturing costs by only monitoring the main uncertainty sources. Composite vehicle components can be optimized considering the impact of uncertainty, yielding a safer yet lighter design.

The Gaussian Process modeling tool developed during the project has been generalized into a user-friendly graphical user interface. The advantage of this tool is twofold; (i) The simplicity yet complimentary user interface allows engineering teams company-wide to benefit from this powerful tool, and (ii) Gaussian process models can reduce the effective simulation turn-around time from days to seconds, enabling the use of these models for uncertainty quantification and propagation purposes as well as design optimization.

Uncertainty sources are generally categorized as aleatory and epistemic. While the former uncertainty source is inherent to the system (and hence irreducible), the latter is generally due to lack of knowledge or data, and may be reduced by conducting more simulations, experiments, or in-depth studies. In the case of materials, both sources are present and may be introduced in the design and constituent selection stages, manufacturing processes, or operation. Such uncertainties manifest as, e.g., mechanical (e.g., Young's modulus, Poisson ratio, yield stress, damage evolution parameters, etc.) or geometrical (e.g., reinforcement distribution, fiber misalignment in composites) variations. To elaborate more on material uncertainty, we take woven fiber composites as an example. These materials have been increasingly used in aerospace, construction, and transportation industries because of their superior properties such as high strength-to-weight ratio, non-corrosive behavior, enhanced dimensional stability, and high impact resistance. Woven fiber composites possess, as illustrated in FIG. 171, a hierarchical structure that spans multiple length-scales from nanoscale to macroscale. Within each of these length-scales, many correlated and spatially varying uncertainty sources are introduced: The high pressure and flow of the resin or draping change the local architecture of the fibers during the preforming process. Additionally, processing variations and material imperfections cause the fiber volume fraction to spatially vary across the sample. These variations are particularly pronounced along the yarn paths where there is compact contact. These macroscopic uncertainties are manifestations of many uncertainty sources that exist at the finer scales where the number and dimensionality of the uncertainty sources increase due to the delicacy of materials.

FIG. 170 shows a multiscale structure: Schematic view of a four-scale woven fiber composite with polymer matrix. In computational modeling of this structure, each integration point at any scale is a realization of a structure at the next finer scale.

Fibrous composites have been previously investigated to determine how much their properties and performance are sensitive to uncertainty. The focus of these works, however, has not been placed on rigorously modeling the uncertainty sources and statistically propagating their effects across multiple scales. For instance, modeling spatial variations via RFs, connecting them across different spatial scales, and investigating stochastic simulations are often neglected. Savvas et al. studied the necessary RVE size as a function of spatial variations of fiber volume fraction and yarn architecture. Their research showed that the RVE size should increase at higher fiber volume fractions. They also concluded that the mesoscale RVE size is more affected by fiber orientation than waviness. Their further studies illustrated that geometrical characteristics (i.e., the shape and arrangement of the fibers) and the material properties (Young's moduli of the constituents) affect the homogenized response in UD composites quite significantly (with the former being more important). The variations were shown to decrease as the number of fibers and RVE size increased. Average axial and shear stiffness constituted the response in these studies. Vanaerschot et al. studied the variability in composite materials' properties and concluded that the stiffness in the mesoscale RVE is affected by the load orientation and, additionally, it significantly decreases as the fiber misalignment increases. Hsiao and Daniel experimentally and theoretically investigated the effect of fiber waviness in UD composites and demonstrated that it decreases composite's stiffness and strength under uniaxial compression loading. Komeili and Milani devised a two-level factorial design at the mesoscale to study the sensitivity of orthogonal woven fabrics to the material properties and yarn geometry. They illustrated that, based on the applied load, these parameters could have a significant effect on the global response (i.e., reaction force). A similar sensitivity study based on Sobol's indices was conducted in to demonstrate that the friction coefficient and yarn height significantly affect the macroscale mechanical response of interest in dry woven fabrics. Yarn properties are spatially homogeneous and there is no fiber misalignment.

To address the shortcomings of the prior works on UQ in woven composites, we employ RFs which are collections of random variables indexed in either time or space. We introduce the Top-down sampling method that builds nested RFs by treating the hyperparameters of one RF as the responses of another RF. This nested structure allows us to model non-stationary and C0 (i.e., continuous but not differentiable) RFs at fine length-scales (i.e., mesoscale and microscale) with a stationary and differentiable RF at the macroscale. We motivate the use of multi-response Gaussian processes (MRGPs) to parsimoniously quantify the RFs and conduct sensitivity analyses for dimensionality reduction. The resulting approach is non-intrusive (in that the computational models need not be adapted to account for the uncertainties) and can leverage statistical techniques (such as metamodeling and dimensionality reduction) to address the considerable computational costs of multiscale simulations.

For UQ of UD composites, we focus on two microstructural uncertainty sources that can be captured by imaging techniques: fiber distribution and fiber waviness. Microstructure images of UD plates are taken at Ford, from which we measure these two quantities of interest (QoI) then model their variation with statistical methods. We address the challenge of simulating a nonstationary and nonhomogeneous RF for fiber distribution modeling by introducing a hierarchical nonparametric statistical model. For fiber waviness, the image data of which is extremely limited, assumptions such as stationarity are made and a time series approach is applied to generate realistic samples from the underlying distribution. The two methods are integrated into a data-driven sampling algorithm that can simulate the spatial distributions of the two QoIs simultaneously for further computational mechanics analysis.

The developed methods and tools are applied to fiber composites which have been increasingly used in aerospace, construction, and transportation industries due to their superior performance. Our contributions, hence, have far reaching impacts on various sectors of the economy.

Multiscale UQ and UP with Application to Cured Woven Composites

Our approach for multiscale UQ and UP has two main stages: Intra-scale UQ and inter-scale UP. We start by identifying the uncertainty sources at each scale and modeling them via RFs where one RF is associated with each structure realization. We employ RFs with sensible (i.e., physically interpretable) parameters for three main reasons: (i) To couple uncertainty sources across length-scales and enable their propagation from lower to higher scales, (ii) To connect the most important parameters of the RFs to the features of the material system and hence identify the dominant uncertainty sources in a physically meaningful way, and (iii) To allow for a non-intrusive UQ procedure by directly using the RFs' outputs in the multiscale FE simulations (instead of adapting the FE formulations for UQ and UP). Due to these reasons, we use the best linear unbiased predictor (BLUP) representation of multi-response Gaussian processes. MRGPs enable sensible characterization of uncertainty sources, are flexible and computationally efficient, and can be easily trained via available data.

At this point, the dimensionality in the UQ and UP tasks has been reduced from the number of degrees of freedom in the multiscale simulation to the few hyperparameters of the MRGP at the coarsest scale. However, depending on the material system and quantities of interest, generally not all the hyperparameters need to be considered in the UP process. Hence, further dimensionality reduction can be achieved by identifying the dominant uncertainty sources and, equivalently, the corresponding RF parameters through, e.g., sensitivity analysis.

The second stage of our approach starts by replacing the nested simulations at fine scales via inexpensive but accurate metamodels (aka surrogates or emulators) to decrease the computational costs of a single multiscale simulation from hours (or even days) to minutes. The choice of the metamodel, its inputs, and its outputs depend on the nature of the FE simulation. Finally, the uncertainty at the highest scale is quantified by propagating the uncertainty from all the finer scales in the UP process. During UP, various multiscale simulations are conducted where for each simulation one realization of the spatially varying quantities are used in the multiscale material. To generate each of these realizations, we introduce the Top-down sampling approach where realizations are assigned to the spatially varying parameters from the coarsest scale to the finest scale in the material system. This sampling strategy enables modeling (i) non-stationary and C0 (i.e., continuous but not differentiable) spatial variations at the fine scales, and (ii) partial correlations between the various uncertainty sources within and across scales. Although the top-down sampling method can be integrated with any analytical RF, we have chosen MRGPs since they are sufficiently flexible and possess a few hyperparameters which are all physically interpretable. Additionally, other RFs can sometimes be converted into GPs upon appropriate transformations. Our approach is demonstrated for a composite with two length-scales in FIG. 172.

FIG. 171 shows demonstration of our approach for s two-scale structure: Spatial random processes (SRPs) are employed for generating spatial variations which are connected through the top-down sampling procedure.

Multi-Response Gaussian Processes for Uncertainty Quantification

MRGPs are widely popular in RF and surrogate modeling and have been used in a wide range of applications including UQ, machine learning, sensitivity analyses of complex computer models, Bayesian optimization, and trac Table 9-Bayesian calibration. For an RF with q outputs y=[y1, yq]T and the field (e.g., spatial or temporal) inputs x=[x1, , xd]T, the BLUP representation of an MRGP with constant prior means reads as:


yaboutq(β,c(x,x′)),  (15-17)

where q represents a q-dimensional Gaussian process, β=[β1, , βq]T is the vector of responses' means, and c(x,x′) is a parametric function that measures the covariance between the responses at x and x′. One common choice for c(x,x′) is:


c(x,x′)=Σ⊗ exp{Σi=1d−10ωi(xi−x′i)2}=Σ⊗r(x,x′),  (15-18)

where Σ is a q×q symmetric positive definite matrix that captures the marginal variances and the covariances between the outputs, d is the dimensionality of the field, ω=[ω1, , ωd]T are the so-called roughness or scale parameters that control the smoothness of the RF, and ⊗ is the kronecker product. Note that the dimension of β and Σ depends on q, while that of ω depends on d. The parameters β, Σ, and ω are called the hyperparameters of an MRGP model and, collectively, enable it to model a wide range of random processes:

    • The mean values of the responses over the entire input space are governed by β.
    • The general correlation between the responses (i.e., yi and yj, i≠j) over the input space is captured by the off-diagonal elements of Σ.
    • The variations around the mean for each of the responses are controlled by the diagonal elements of Σ.
    • The smooth/rapid changes of the responses across the input space are controlled by ω.

In case some experimental data are available, all the hyperparameters of an MRGP model can be estimated via, e.g., the maximum likelihood method. Otherwise, as in this work, expert or prior knowledge can be used to adjust these parameters and model a spatially varying quantity. Once these hyperparameters are determined, generating realizations from an MRGP model is achieved through a closed-form formula.

Top-Down Sampling for Uncertainty Propagation

To carry out one multiscale simulation, material properties must be assigned to all the IPs at all scales where the IP properties at any scale depend on an RVE at the next finer scale (this RVE itself has many IPs). Since these properties depend on the uncertainty sources (or, equivalently, on the RFs), the latter must be coupled across the scales. Recall that, due to the multiscale nature of the structure, the number of RFs significantly increases at the fine scales because we associate an RF to each structure realization.

Having used RFs whose parameters are physically sensible and can be directly linked to the uncertainty sources, this cross-scale coupling is straightforward and can be achieved with top-down sampling where the outputs of the MRGP at each IP of a particular scale serve as the hyperparameters of the MRGP of the RVE associated with that IP. This process constitutes nested RFs. To assign values to the IP parameters in the entire multiscale structure, this approach starts from the coarsest or top scale and hence the name top-down sampling.

While the Top-down sampling method works with any parametric RF representation (e.g., PCE or KL expansion), we highly recommend employing compact representations that include a few hyperparameters. This is mainly because the number of hyperparameters at the coarse scales increases rapidly as the number of spatially varying quantities increases at the fine scales. For instance, assuming three (two) quantities change spatially in a 3D microstructure, an MRGP with 12 (8) hyperparameters is required. To model the spatial variations of these 12 (8) hyperparameters at the mesoscale, an MRGP with 93 (47) hyperparameters is required.

Case Study on Cured Woven Fiber Composites

We now follow the steps of our approach to quantify the macroscale uncertainty in the elastic response of a cured woven composite as a function of spatial variations in seven uncertainty sources: fiber volume fraction and misalignment, matrix and fiber modulus, and yarns' geometry parameters (i.e., yarn angle, height, and spacing). As illustrated in panel (a) of FIG. 172, the structure is composed of four identical woven plies that are stacked in the same orientation and constitute a total thickness of 2.4 mm (the fiber orientations are indicated with light blue lines).

FIG. 172 shows the macroscopic cured woven laminate structure. Panel (a) shows the deformed structure. The light blue lines indicate the fiber orientation. The dimensions are scaled for a clearer representation. Panel (b) shows the deterministic spatial variations of yarn angle obtained from simulating a perfectly manufactured composite. Panel (c) shows Von Mises stress field corresponding to Case 9. Panel (d) shows the random spatial variations of yarn angle corresponding to one of the realizations of Case 1. Panel (e) shows the random spatial variations of θ1 corresponding to one of the realizations of Case 3.

The geometry of these woven plies is obtained via the bias-extension simulation of woven prepregs using the non-orthogonal constitutive preforming model. While the bottom of the sample is clamped, the other end is pulled by 1 mm to generate the bias tension deformation. In the macroscale simulations, 3D solid continuum elements are employed to discretize the structure. As our focus is on UQ and UP, at this point we have assumed that only elastic deformation occurs.

Uncertainty Sources

Longitudinal fiber and matrix moduli, Ef and Em, are the first two uncertainty sources. Given the moduli, the yarn material properties primarily depend on two parameters: fiber volume fraction (in yarn), v, and fiber misalignment. While in most previous works v is postulated to be spatially constant, in practice, it varies along the yarn path particularly where yarns have compact contact. Consequently, our next uncertainty source arises from the spatial variations of v which starts from the microscale and propagates to mesoscale and macroscale. In this work, we have assumed that 45%≤v≤65% based on our material system.

During the manufacturing process, the fibers in the yarn deviate from the ideal orientation and render the cross-section of the yarn heterogeneous and anisotropic. These deviations result in fiber misalignment which is different from the concept of fiber waviness in that a fiber can be perfectly waved without misalignment. As illustrated in FIG. 173, this misalignment can be characterized by the two angles ϕ and θ which measure the deviation of the fiber direction, ƒ1, with respect to the local orthogonal frame on the yarn cross-section, gk. Based on the available experimental data in the literature, in this work we have assumed −π≤φ≤π and 0°≤θ≤90°.

In modeling the mesoscale woven composites, the yarn architecture is often presumed to be perfect where the yarn angle, α, is set to 90° and the yarn height, h, and spacing, s, are fixed to their nominal values. These assumptions do not hold in practice due to the large in-plane shear deformation during preforming process and manufacturing imperfections. Hence, we also investigate the effect of the spatial variations of the woven RVE architecture (α, h, and s) on the macroscopic properties.

FIG. 173 shows fiber misalignment angles. The zenith and azimuth angles characterize the fiber misalignment angle with respect to the local orthogonal frame on the yarn cross-section.

Lastly, we note that in our example the deterministic spatial variation of α in a perfectly manufactured composite is, as opposed to the other parameters (i.e., [v, φ, θ, h, s]), available from the preforming process simulation. This deterministic variation is used as the prior mean (β in Eq. (9-17)) of spatial distribution of α while for the other six parameters the nominal values are employed as the (spatially constant) prior mean. In all seven parameters, the posterior spatial variations are stochastic.

We employ the computational homogenization technique for modeling the multiscale woven sample where the material property at any length-scale is calculated through the homogenization of an RVE at the lower scale. At the microscale, the RVEs include 300×300×60 voxels (42 μm×42 μm×8.4 μm) and the fibers have a diameter of 7 μm. The simulations are elastic where periodic boundary conditions (PBCs) are employed. It is assumed that the fibers and the matrix are well bonded and there are no voids. To obtain the stiffness matrix, C, of the UD RVE, six stress-free loading states are applied (i.e., only one of the εxx, εyy, szz, εxy, εxz, and εyz strain components are applied in each case). Since the simulations are elastic, C mainly depends on the volume fraction, v.

At the mesoscale, the open source software TexGen is used to create the geometry and mesh for the 2×2 twill woven RVE with 8 yarns. The space between the yarns is filled with matrix and voxel meshes are used to discretize the RVE where each voxel is designated to either a yarn or the matrix. To balance cost and accuracy, we have used a voxel mesh with 625000 elements. To reduce the computation costs, PBCs are employed throughout.

The nominal properties of carbon fibers and epoxy resin were taken from manufacturer's data (see Table 9-14). The resin is isotropic, and its material properties are taken from pure epoxy. Yarns with well-aligned fibers are treated as transversely isotropic. With fiber misalignment, however, yarns are not transversely isotropic since the material frame across the IPs on their cross-section is non-uniformly distributed. In this case, the micro-plane triad model is employed to account for fiber misalignment by defining an orthotropic micro-triad, {right arrow over (ƒ)}k, for each IP of the yarn. This triad is related to the local frame, {right arrow over (g)}k (see FIG. 174), via the misalignment angles:

f 1 = cos ( θ ) g 1 | g 1 | + sin ( θ ) cos ( φ ) g 2 | g 2 | + sin ( θ ) sin ( φ ) g 3 | g 3 | , ( 9-19) f 2 = - sin ( θ ) g 1 | g 1 | + cos ( θ ) cos ( φ ) g 2 | g 2 | + cos ( θ ) sin ( φ ) g 3 | g 3 | , ( 15-20) f 3 = f 1 ^ f 2 , ( 15-21)

where |⋅| and {circumflex over ( )} denote, respectively, the norm of a vector and the cross product. As for the local frame {right arrow over (g)}k, it is automatically generated by TexGen for each IP (each voxel at the mesoscale) once the woven RVE is discretized. We note that, the stiffness matrix at each yarn material point is obtained via the UD-RVE homogenization.

To link the mesoscale and macroscale, the stress-strain relations for effective elastic material properties of woven RVE are required. This relation can be written in terms of the symmetric mesoscale stiffness matrix as:

[ σ 1 1 σ 2 2 σ 3 3 σ 1 2 σ 1 3 σ 2 3 ] = [ C 11 C 12 C 13 C 14 0 0 C 22 C 23 C 24 0 0 C 33 C 34 0 0 C 44 0 0 Sym . C 55 C 56 C 66 ] [ ɛ 1 1 ɛ 2 2 ɛ 3 3 2 ɛ 1 2 2 ɛ 1 3 2 ɛ 2 3 ] , (15-22)

TABLE 9-14 Fiber and matrix properties: The moduli (i.e., E and G) are all in GPa. Poisson's ratios along different directions are also provided. Ezz Exx = Eyy vzx = vzy vxy Gxz = Gyz Gxy Carbon 275 19.8 0.28 0.32 29.2 5.92 fiber Epoxy 3.25 3.79 0.39 0.39 1.36 1.36 resin

Top-Down Sampling, Coupling, and Random Field Modeling of Uncertainty Sources

To help clarify the descriptions, we first introduce some notation. We denote the three scales with numbers: 1→Macro, 2→Meso, 3→Micro.Superscript and subscripts denote, respectively, scales and IPs. Variables with a bar represent averaged quantities over all the IPs at a particular scale. For instance, vi1 denotes the fiber volume fraction assigned to the ith IP at the macroscale.

θ ¯ 2 = 1 N Σ i = 1 N θ i 2

represents the average misalignment (zenith) angle at the mesoscale for a woven RVE.

The uncertainty sources in our composite are summarized in Table 9-15. While some sources are only defined among different structures (under spatial variations across realizations), others possess an extra degree of variation in that they also change within structures.

TABLE 9-15 Uncertainty sources for one macro structure realization: The sources are fiber misalignment angles (θ and φ), yarn spacing and height (s and h), fiber and matrix moduli (Ef and Em), fiber volume fraction (v), and yarn angle (α). Scale Uncertainty Sources n Micro v3, Ef, Em None Meso θi2, φi2, vi2, Ef, Em, α2, s2, h2 θi2, φi2, vi2 Macro NA θi1, φi1, vi1, Efi1, Emi1, αi1, si1, hi1

Assuming the eight tows in a woven RVE are statistically independent and the spatial variations within them originate from the same underlying random process, a total of 12 hyperparameters are required to completely characterize the spatial variations of [θi2, φi2, vi2] by an MRGP (see Eq. (9-17)): three mean values (β=[βv, βφ, βθ]T), six variance/covariance values for Σ([σvv2, σφφ2, σθθ2, σ2, σ2, σφθ2]), and three roughness parameters (ω=[ωx, ωy, φz]T where xyz denotes the cartesian coordinate system at the mesoscale). Once these parameters are specified, the spatial coordinates of the IPs in a woven RVE can be used to assign realizations of v, φ, and θ to them. For each IP at the macroscale, however, these 12 hyperparameters serve as some of the responses of the macroscale MRGP whose other responses correspond to [Efi2, Emi2, αi2, si2, hi1]. Therefore, the macroscale MRGP has a total of 173 hyperparameters (17 mean values for β, 153 unique covariance/variance values for Σ, and 3 values for ω=[ωη, ωξ, ωζ]T where ηξζ denotes the cartesian coordinate system at the macroscale). In the top-down sampling approach, first the 173 hyperparameters of the macroscopic MRGP are prescribed. Then, this MRGP is sampled to assign 17 values to each macroscopic IP where 12 of them serve as the hyperparameters of the mesoscopic MRGPs, 3 of them are employed to construct the woven RVE, and 2 of them to determine the fiber and matrix moduli. This process is illustrated in FIG. 174 where, for clarity, only two (out of the 17) hyperparameters at the mesoscale are presented. FIG. 174 shows coupling the uncertainty sources across the scales: The spatial variations of v and θ at the macroscale are connected to those at the finer scales. For brevity, the coupling is illustrated only for the average values for the two quantities (i.e., the mean of the RFs: β=[βv, βθ]=[v2,θ2]).

Dimension Reduction at the Mesoscale Via Sensitivity Analysis

By modeling the spatial variations via RFs, the dimensionality of the UQ and UP problem has decreased to the number of RF hyperparameters. Although this is a significant reduction, the considerable cost of multiscale simulations (even in the linear analysis) renders the UQ and UP process computationally demanding. To address this issue, we note that depending on the property of interest a subset of uncertainty sources are generally the dominant ones in physical systems. Since our composite undergoes an elastic deformation, we expect a small subset of the uncertainty sources (i.e., RF hyperparameters) to be important.

We conducted multiscale sensitivity analyses to determine which of the 12 hyperparameters of an MRGP model are the most important ones (and must be considered in UP) based on their impact on mesoscale material response. Our studies included changing one of the hyperparameters (while keeping the rest of them fixed) and conducting 20 simulations to account for the randomness. Then, if the variations in the homogenized response (effective moduli) were negligible, the hyperparameter was deemed as insignificant and set to a constant thereafter. All the simulations in this section were conducted on a woven RVE with α=90°.

TABLE 9-16 Case studies to determine the effect of fiber misalignment and its spatial variations: The triplets in the description fields correspond to (θ2, var(θ2), var(φ2)). Case ID Description 1 (1°, 1, 202) 2 (1°, 5, 202) 3 (1°, 10, 202) 4 (3°, 1, 202) 5 (3°, 5, 202) 6 (3°, 10, 202) 7 (5°, 1, 202) 8 (5°, 7, 202) 9 (5°, 10, 202) 10 (1°, 1, 502) 11 (1°, 5, 502) 12 (1°, 10, 502) 13 (3°, 1, 502) 14 (3°, 5, 502) 15 (3°, 10, 502) 16 (5°, 1, 502) 17 (5°, 7, 502) 18 (5°, 10, 502)

We found that the homogenized moduli are affected by neither the six covariance/variance values (i.e., [σvv2, σφφ2, σθθ2, σ2, σ2, σφθ2]) nor the three roughness parameters ω=[ωx, ωy, ωz]T. In case of average values (i.e., β), the average fiber volume fraction (v) and zenith angle (θ), as opposed to that of the azimuth angle, were found to considerably affect the homogenized response of a woven RVE at the mesoscale. The effect of average values on the effective moduli are summarized in FIG. 175. As illustrated in panel (a) of FIG. 175, the moduli increase linearly as a function of average fiber volume fraction. Comparing the first nine Cases in panel (b) of FIG. 175 with the last nine Cases, it can be concluded that var(φ2) insignificantly affects the moduli and poison ratios. panel (b) of FIG. 175 also indicates that the deviations from the reference solution increase as θ2 increases (θ2 increases between the jumps, see Table 9-16).

FIG. 175 shows effect of average values on the effective moduli of a woven RVE: panel (a) Effect of fiber volume fraction and, panel (b) Effect of misalignment. Each point on these figures indicates the average value over 20 simulations. The Case IDs in panel (b) are defined in Table 9-15. The reference solution refers to a case where there is no misalignment.

It is noted that since we are interested in the elastic response of the multiscale composite in FIG. 173, the variations of effective moduli were only considered in our sensitivity studies. As opposed to the effective behavior, the local behavior (i.e., stress field) of the woven RVE is quite sensitive to the spatial variations of both v and θ (but not φ) and must be considered in nonlinear analysis.

Replacing Meso and Microscale Simulations Via Metamodels

To further reduce the multiscale UQ and UP costs, we employ metamodels to replace the micro and mesoscale FE simulations corresponding to each macroscale IP. In particular, the metamodel captures the macroscale spatial variations of the stiffness matrix of the woven RVEs associated with the macroscale IP's as a function of yarn angle (α2), average volume fraction (v2), yarn height (h) and spacing (s), average misalignment angle (θ2), and fiber and matrix moduli (Ef and Em). In machine learning parlance, the inputs and outputs of the metamodel are, respectively, [α2, v2, s2, h2, θ2, Ef, Em] and the stiffness matrix C of a woven RVE. To fit the metamodel, we generated six training datasets of sizes 60, 70, , 110 with Sobol sequence for α2∈[45°, 90°], s2∈[2.2, 2.5] mm, h2∈[0.3, 0.34] mm, v2∈[40%, 70%], Ef2∈[150, 400] GPa, Em2∈[1.5, 5] GPa, and θ2∈[0°, 6° ]. Afterwards, we fitted an MRGP metamodel to each dataset. The accuracy of each model was then evaluated against a validation dataset with 30 samples via:

e = 1 0 0 1 3 0 Σ i = 1 3 0 ( 1 - y ^ i y i ) 2 % , (15-23)

where ŷ=[ŷ1, , ŷ30] and y=[y1, , y30] are obtained from, respectively, the metamodel and FE simulations. The prediction error of each model is illustrated in FIG. 176 where it is evident that with roughly 100 samples all the elements of C can be predicted with less than 6% error.

FIG. 176 shows prediction error as a function of the number of training samples: As the number of training samples increases, the accuracy of the MRGP metamodel in predicting the elements of the stiffness matrix of the mesoscale woven RVE increases.

Graphical User Interface (GUI) for Optimal Sampling and Metamodeling

As metamodeling is a broadly applicable tool (also outside the field of stochastic multiscale modeling), two user-friendly Graphical Interfaces have been developed: Optimal Latin Hypercube Sampling (OLHS), and Gaussian Process modeling. The ideology behind these tools is to be functional and complete, while being intuitive enough for novice users. Furthermore, the graphical interfaces have been developed using Matlab Guide and can be run on any 64-bit computer under the windows operating environment.

As shown in panel (a) of FIG. 178, OLHS Generation and visualization package allows for the automatic generation of training data inputs that span the desired metamodel space as optimal and uniform as possible. Any number of input variables and number of samples can be specified. As shown in panel (b) of FIG. 178, our package for GRP is applicable to multi-dimensional and multi-response datasets and can automatically handle noisy observations once enough training data is provided. Additionally, the interface includes features that allow the user to evaluate metamodel accuracy, perform prediction on unobserved inputs, and visualization that allows the user to conveniently investigate input-output relations regardless of problem dimensionality.

Results on Macroscale Uncertainty

Multiple macro simulations are conducted where θi1, vi1, Efi1, Emi1, αi1, si1, and hi1 change spatially across the macroscale IPs. To quantify the importance of these variables' spatial variations on the macroscopic behavior, eight cases are considered. The spatial variations are changed with an MRGP at the macroscale in a controlled manner from one case to the next. Except for the last case, 20 independent macroscale simulations are conducted for each case to account for the randomness. In summary, a total of 161 macroscale simulations are conducted. The last case study serves as the reference where there is no misalignment, αi1 is determined via the processing simulation over the structure as illustrated in panel (b) of FIG. 172, and all other parameters are set to their nominal values, i.e., vi1=55%, Efi1=275 GPa, Emi1=3.25 GPa, si1=2.35 mm, and hi1=0.32 mm. In cases 1 through 7, the effect of spatial variations is considered for one parameter at a time where the amount of variations with respect to the prior mean is controlled by the variance of the macroscale MRGP. In these cases, we set the variance associated with the spatially varying parameter to, respectively, 9, 25, 1, 0.052, 0.012, 402, and 1. These variances are large enough to capture realistic spatial variations while small enough to ensure that the realized values are not outside the ranges where the mesoscale MRGP metamodel is fitted. Sample spatial variations for case 1 and 3 are demonstrated in panels (b)-(e) of FIG. 172.

Since in reality the uncertainty sources coexist, in case 8 we consider the effect of all the uncertainty sources and their correlation. Here, the individual variances (diagonals of macroscale MRGP, Σ) are the same as in cases 1 through 7 while the covariances are chosen to reflect the negative correlation between the fiber volume fraction and both yarn and fiber misalignment. To this end, we choose [σαv, σαθ, σ]=[−9, 3, −3] and set the rest of the off-diagonals of Σ to zero. We note that σαv is negative to model the increase in fiber volume fraction as the yarns get closer after preforming. σ is also negative to consider the decrease in misalignment angle in rich fiber regions.

TABLE 9-17 Description of the simulation settings: Nine simulation cases are considered to quantify the macroscale uncertainties and the relative importance of spatial variations. MRGPs are employed to generate random realizations in all cases. Except for case 9 (reference simulation), all cases include 20 simulations to account for randomness where only one uncertainty source exists. Description Case 1 Spatial variations in α with its prior spatial distribution determined by processing simulation Case 2 Spatial variations in fiber volume fraction with a constant prior of 55% over the structure Case 3 Spatial variations in fiber misalignment with a constant prior of 3° over the structure Case 4 Spatial variations in yarn spacing with a constant prior of 2.35 mm over the structure Case 5 Spatial variations in yarn height with a constant prior of 0.32 mm over the structure Case 6 Spatial variations in fiber modulus with a constant prior of 275 GPa over the structure Case 7 Spatial variations in matrix modulus with a constant prior of 3.25 GPa over the structure Case 8 Spatial variations in all parameters Case 9 All parameters set to nominal values (α from processing simulation) and no misalignment

Panel (a) of FIG. 178 compares the exerted force on the cross-section of the sample for the nine cases where, for cases 1 through 8, the force-displacement line is averaged over the 20 multiscale simulations. The results suggest that in the presence of fiber misalignment, the structure weakens and hence the reaction force decreases. This weakening is exacerbated in the presence of other uncertainty sources. In particular, while in the reference case the maximum reaction force (magnitude) is 15.4 kN, in cases 3 and 8 it is 14 kN and 13.5 kN, respectively. These results indicate that: (i) Fiber misalignment must be accounted for in the simulation of composites even if one is primarily interested in the global response in linear analyses. (ii) The interaction among various uncertainty sources should be considered: once all the uncertainty sources are accounted for, the structure is more noticeably weakened with a 12.3% reduction in reaction force. (iii) The insensitivity to the spatial variations of some parameters can be explained by the fact that in linear analyses the global response is mainly affected by the averaged properties. Since we only introduce spatial variations (we do not change the parameter averages over the structure except for fiber misalignment), the results intuitively make sense.

To illustrate the effect of spatial variations on local behavior, we compare the average and standard deviation of the von-Mises stress field over the mid-section of the structure in panels (b)-(c) of FIG. 178, respectively. We choose the mid-section over the entire structure for analyzing the strain field since an explicit solver is employed in our multiscale FE simulations where artificial stress concentrations might occur at the boundaries of the structure. The curves in these two figures are obtained by analyzing the 20 simulations corresponding to each case. Similar to panel (a) of FIG. 179, panel (b) of FIG. 178 demonstrates that the structure weakens in the presence of fiber misalignment as the mean stress over the mid-section is lower than that of the reference structure (case 9). It is evident that the most realistic case where all the uncertainty sources are present results in the weakest structure. More interestingly, this weakening is not uniform over the mid-section and is the largest at the middle where the reduction is roughly 12.3%.

The highest variations among the simulations of a specific case are observed in case 8, where all the parameters change spatially and relatively, see panel (c) of FIG. 178. Cases 2, 3, and 7 are next in line with cases 2 and 7 having more fluctuations among the IPs (note that each point along the x-axis corresponds to an IP in the mid-section). Finally, cases 4 and 5, which correspond to the simulations where the yarn spacing and height change, have the least amount of variations. This is expected since the stiffness matrix of a woven RVE (obtained via the mesoscale MRGP metamodel) is also insignificantly sensitive to these parameters. We highlight, however, that we have considered a relatively small range of variations for these two parameters. Their effect will be more prominent if these ranges are increased.

Image-Based Microstructural UQ of UD Composites

The UQ work for UD composites aims at statistical modeling and reconstruction of the material microstructural features, including (1) non-uniform fiber spatial distribution and (2) fiber waviness. We build our models based on microstructure images of UD plates taken at Ford Motor Company. Several machine-learning- and applied-statistics-based approaches are developed for image characterization, information retrieval and generative model building. For the spatial fiber distribution, which exhibits both non-stationarity and non-homogeneity, we first model the data (image) via a tree-regression algorithm then a hierarchical nonparametric sampling approach is developed. The approach is completely data-driven, in the sense that no probability models are assumed and a part of a new sample is generated by resampling from the data.

Fiber waviness is the local orientation of the fiber bundles relative to the global direction of the fibers. Perfectly straight fibers have zero waviness everywhere, however, the transverse images taken from unidirectional fiber composite samples (FIG. 182) show that waviness does exist. On the other hand, due to the limitation in image quality, only partial fibers are observed in terms of disconnected fragments therefore traditional computer vision algorithms for edge or object detection are not applicable in the invention. To conquer this challenge, we developed a segmented regression algorithm that can estimate the local waviness angle via a linear-regression-like-, optimization-based approach. Then the angle distribution along the fiber longitudinal direction are modeled with a time series statistical model, from which we can sample realizations from.

Fiber Distribution Modeling

By visually inspecting the distribution of fibers, which is represented by distribution of local volume fraction of fibers (panel (a) of FIG. 179), the spatial correlation between the local fiber volume fractions exhibits two characteristics. The first one is non-homogeneity, which is reflected by the layered structure along the vertical direction. The second one is non-stationarity, which refers to the local curvature of the volume fraction patterns along the horizontal direction. Modeling both features with traditional parametric statistical models will involve estimation of lots of parameters representing these features. Therefore, we use an alternative approach: along the vertical direction, the non-homogeneity, i.e., the shape of some wave-like functions, is sampled from its empirical distribution through resampling from the image dataset (with replacement); along the horizontal direction, the non-stationarity is modeled by looking for a representative statistic of this feature and attempting to find the distribution of the statistics.

The first step to find such a statistic is to reduce the dimension of each sample image from approximately 450,000 pixels to a manageable size. To compress the data in a sample, we used a regression tree algorithm, i.e., a sample is represented by a tree-structured field. The locations of the splitting lines (nodes of the tree) are found by minimizing the integrated relative error (IRE), and between the separation lines the data are interpolated linearly. By setting the regression goal to IRE <5%, we normally obtain 200-300 nodes. Since the splitting lines separate the most distinct areas, the locations of them contain the information of non-stationarity: the denser the lines are, the more local curvatures the area include (panel (b) of FIG. 179). Note that a dense collection of the lines means smaller inter-line distances. It follows that we can use the distribution of the inter-line distances to characterize the non-stationarity.

The distribution, in terms of probability density estimates, of the inter-line distances are generated with all the samples. The first three plots of FIG. 180 shows three of them. It can be observed that each sample has different levels of non-stationarity, which might account for the difference in the material performance. The distributions in the last three plots are estimated by grouping inter-line distance data of 10 randomly selected samples together. The similarity in shape of the probability densities suggests the convergence of the distribution, i.e., the common distribution behind all samples.

This observation provides a way of generating random samples with similar levels of non-stationarity, which is sampling the locations of the splitting lines from the common distribution in panel (b) of FIG. 180. The volume fractions along the vertical splitting lines are sampled directly from the empirical distribution estimated from the data. Therefore, a sampling algorithm for the fiber distribution/volume fraction can be constructed by (a) generating random splitting locations from the common distribution of inter-line distances, then (b) resampling the volume fractions along the splitting lines from the dataset, and (c) the areas between the lines are filled by interpolating the neighboring lines. Two example reconstructions are demonstrated below (FIG. 181), which shows both randomness and high level of similarity compared to the original samples (for example, in panel (a) of FIG. 180).

Fiber Waviness Characterization

Characterization of fiber waviness from the transverse images is not as straightforward as the same task with fiber distribution, as the local curvature in the image, in terms of local slopes or angles, cannot be calculated directly from the pixel information (e.g., binary or grayscale values). Normally the characterization process would involve detecting each of the fibers on the image and calculate the angles accordingly. However, this approach fails in the invention because the fibers in the images are not fully shown, i.e., only partial fibers are observed, and some of the parts appear in just dots or small pieces, the orientation of which are not measurable individually (see, for example, the binary image segment in FIG. 182). It is possible to design filters to remove them as noises, but the fact that they still have a general trend as a group indicates that statistical characterization is helpful in preserving the information. Also note that linear regression is an ideal tool to estimate slopes, therefore we developed a segmented linear regression approach (FIG. 182) to characterize the local curvature in the images.

The idea behind this method is: the relationship between the local angle and local slope is given by β=tan(α), where β is the slope and α is the angle, and the local slope is estimated by the slope of the regression line with the points on a locally binarized segment of the original image. The challenging part is to build a valid linear model for the regression: the classic simple linear model assumes the error is normally distributed and the well-known least squares method is derived based on this assumption, which is not valid in our problem. We proposed our own regression algorithm customized for this case: under the assumption that the fiber pixel point and fiber locations are uniformly distributed, assume the origin of the coordinates is put in the center of the image segment, and a regression-through-the-origin mean prediction is given by the function y=βx, the estimate of the slope β is given by:

β ^ = argmax β f ( e ; β ) (9-24)

where e is the set of residuals with the regression line y=βx, ƒ:→ (assume there are n points in total) is the mapping between the residuals and the number of modes in the probability density estimation of the residuals. With this estimation, the trend of the fibers in the transverse direction are correctly captured (see FIGS. 183 and 184 for illustration).

Fiber Waviness Modeling

The challenge associated with this task is the very limited number of images: effectively only one in total (FIG. 184). Therefore, assumptions are made to validate our approach to build the statistical model and sampling algorithm: (1) the waviness contained in this image is representative of the waviness distribution, (2) the waviness distribution is stationary, and (3) the impact of the waviness change along the thickness/vertical direction is trivial, therefore only the fiber angle distribution along the horizontal direction is modeled. Based on (9-3), we can average the waviness angles along the thickness direction and obtain a 1D signal (FIG. 185).

The resulting signal has a wave shape with varying amplitudes and frequencies. A natural characterization of such signal is the periodogram. A periodogram is the estimate of the spectral density of a signal, which can be obtained by discrete Fourier transform of a time or spatial series. Under assumption (2), it can be shown that the periodogram of a series converges in distribution to a sequence of independently and exponentially distributed random variables as the length of the series increases. The sampling algorithm is then constructed by (a) obtaining the periodogram of the signal, (b) generating a random periodogram by sampling from independent exponential distributions with parameters given by (a), and (c) using inverse Fourier transform to obtain the new waviness sample from (b). FIG. 185 shows the comparison between the original sample and the reconstructions in both spatial and frequency domain.

Joint Sampling of Two Uncertainty Sources

The two algorithms above, one for fiber waviness and one for fiber distribution, are capable of modeling and sampling the respective QoI individually. However, when they are integrated into a joint sampling program, spatial constraints must be taken into account. For example, if we want to generate the spatial distributions of the two QoIs for a coupon simulation model simultaneously, the pattern of the fiber distribution at different cross-sections (i.e., cross-sections taken at different locations in the fiber longitudinal direction) should be similar but different, because the fibers are curved along the longitudinal direction. Therefore, combining independent realizations of the two sampling codes will not represent the real situation. In observation of this phenomenon, we developed a joint sampling algorithm for coupon models and realistic realizations can be generated from this method.

Contrary to prior work that rely on random variables, we employ random fields to model the spatially varying uncertainty sources in multiscale materials such as cured woven composites. We introduce the Top-down sampling method that builds nested random fields and, in turn, allows us to model non-stationary variations at fine length-scales (i.e., mesoscale and microscale). We motivate the use of multi-response Gaussian processes to parsimoniously quantify the random fields and conduct sensitivity analyses for dimensionality reduction. The resulting approach is non-intrusive and can leverage statistical techniques to address the considerable computational costs of multiscale simulations.

The computational demand of multiscale materials has been circumvented by the use of Gaussian random processes trained on a space filling design. As computationally demanding simulations are ubiquitous among engineering problems, the developed user-friendly GUIs enable more engineers working on a wide variety of challenges to benefit from these powerful tools. The GUIs and the tools embedded in their source code are able to reduce effective simulation turn-around time from days to seconds, including high dimensional problems of up to 50 input variables.

Image characterization techniques are developed to quantify the variations in UD uncertainty sources (fiber waviness and spatial distribution) and the corresponding statistical models are introduced to study the variability and sample realizations of random fields or random processes from the underlying distribution. Compared to existing work that often pre-assumes some parametric model to represent the uncertainty, our methodology fully utilizes the available microstructural image data and enables the systematic study of the uncertainties of UD composites in a real-world setting.

Uncertainty quantification involves the use of a wide range of computational and statistical tools and the choice of the appropriate tool depends on the available information and resources. For example, for UQ of woven composites, in the absence of data, parametric RF models like Gaussian RFs are chosen and the associated top-down sampling approach for multiscale analysis is developed to study the impact of uncertainty; while for UD of UD composites, images are given, hence the sampling algorithm is designed to mimic the information contained in the image data so that the random reconstructed samples are realistic.

The domain size in UQ is also very important. For instance, the chosen microscale and mesoscale RVEs in UQ of woven composites were sufficiently large and so more uncertainty would have been observed in the response had we chosen smaller RVEs. Finally, we note that the underlying assumptions of any method (including ours) can be validated via experimental data. For instance, our choice of MRGPs for UQ in woven composites implies a normality assumption which might not hold in other applications where the distributions are heavily skewed or appropriate transformations are not readily available. In such cases, other random field representations can be integrated with the top-down sampling approach at the expense of more computations

In the future, we plan to gather experimental images from various structures at different spatial scales to further demonstrate the applications of our approach. Experimental data would additionally allow us to validate the normality assumption made on the marginal variations. Finally, since uncertainty is more important in nonlinear analyses of materials and structures, we plan to apply our approach to nonlinear problems such as plasticity.

For UQ of UD composites, after the variations in the uncertainty sources are quantified and corresponding sampling algorithms are developed, the next step is to conduct FEA simulations based on the generated samples and study the effect of the uncertainty sources. If they are proved to be influential to the properties of the UD composites, other multiscale models that utilize the UD properties should be modified to reflect the presence of uncertainties at the scale of UD. It will also be beneficial if some non-image-based, parametric models are developed for UQ of UD so that the level of uncertainty can be controlled by some parameters to make possible studies like sensitivity analysis and metamodeling.

5. Part-Level Molding and Model Validation

Part-scale preforming experimental validation in this section measures the prediction accuracy of the non-orthogonal prepreg material model and the multiscale preforming simulation method. It also provides guidance about the limitations of current simulation methods and possible solutions in the future. Finally, selection of various process parameters in preforming validation experiments leads to different part qualities. Observation and summary of the relation between process parameters and parts' quality can serve as an empirical rule to produce high-quality parts either for research or for real production application.

With the introduction of double curvature features, the double-dome validation experiment applied in this CIME project gives a trustworthy approach to quantitatively measure prediction accuracy of preforming models for 3D geometry parts.

This double-dome preforming experimental validation can be used for not only the models developed in the invention, but also future models for carbon fiber prepreg preforming simulation to determine their accuracy and application potential.

Double-dome benchmark geometry is used to validate the non-orthogonal material model and the multiscale preforming simulation method developed in the invention at part-scale. Double-dome geometry applied in the invention has 3D shape and complex double curvature features at the size of common automobile parts. It is an ideal benchmark to quantitatively measure prediction accuracy of the preforming models for real part production. In validation, not only the final part shape, but also the yarn angle at different locations, and forming force, are compared. This is because yarn angle, which is an indicator of fiber orientation, significantly affected mechanical stiffness and strength of CFRP parts, while forming force is an indication of membrane stress that controls tow separation and breakage. Different process parameter combinations are tested in this benchmark validation in order to ensure that the models can work at various production conditions. Comparison between the non-orthogonal model and conventional orthotropic material models is also performed to show the improvement we achieved for preforming simulation in the invention.

The major technical target for this part-scale double-dome preforming experiment is to validate the prediction capability of the non-orthogonal material model and the multiscale preforming simulation method we developed in the invention, and to check whether the fiber orientation prediction from these two simulation approaches achieves the proposed 5% error. With successful establishment of this benchmark test and corresponding quantitative measure and validation criteria, we also aim to provide a widely accepted preforming simulation validation method for both academic and industrial researchers.

Preforming is a temperature varying process because of the utilization of hot prepreg sheets and cold/warm tools in the process. In the double-dome benchmark preforming validation experiments performed for the invention, supplied prepregs were first heated in an oven to around 70° C. and then placed in a press for preforming. The geometry of the double-dome punch and the binder are demonstrated in FIG. 186, while the experimental setup is illustrated in FIG. 188. For fast production rate, the press was kept at 23° C. by the coolant running within it, therefore the temperature of the prepregs dropped from the initial value during preforming. In a single layer preforming setup, temperature history at the top surface center, bottom surface center, and one side point on the top surface of one prepreg are measured by thermocouples and plotted in panel (b) FIG. 186. The plot indicates that the prepreg reached a temperature of around 70° C. in the oven. Then, it was cooled down gradually to around 45° C. by air during the transportation from the oven to the press. When placed in the press, the cooling rate increased greatly due to heat conduction between the hot prepreg and cold metal. Specifically, the temperature dropped 20° C. within the first 2 seconds. Meanwhile, it took the press 10 seconds to contact the punch and the prepreg and another 6 seconds to finish preforming. Therefore, the actual temperature of the prepreg during this preforming process was very close to 23° C., i.e., the press temperature. As a result, although the preforming models are capable of taking temperature-dependence into account, in the current validation experimental setup, it is reasonable to characterize the mechanical properties of prepregs and simulate the preforming process at a fixed temperature of 23° C.

Preforming simulation models, utilizing either experiment-based non-orthogonal material model or multiscale method, were established in LS-DYNA® using the dynamic explicit integration method. The simulation setup is illustrated FIG. 188: layers of prepregs are preformed in this process where punch displacement is 90 mm, and binder force increases linearly from 4000 N to 8200 N from experimental measurements. Thickness of the prepregs is orders of magnitude smaller compared to its length and width, so the prepregs are discretized by reduced integrated shell element S4R to reduce the computational cost. Each element is about 4 mm×4 mm with five through-thickness integration points. Tools are simulated as rigid bodies because of their high stiffness compared to the soft prepregs with uncured resin, hence, their element type will not affect simulation results. S3 elements are selected to discretize the tools because of their excellent auto-mesh capability for complex geometries. Friction coefficient between tools and prepregs is set to 0.3 according to the experimental measurement. The friction coefficient is the constant dynamic one in LS-DYNA® and the static friction is neglected because the preforming process leads to large prepreg deformation which, in turn, results in large sliding between the tools and prepregs.

For the first set of validation, only one layer of prepreg with ±45° initial fiber orientation was preformed. The results from the experiment, simulation with experiment-based non-orthogonal material model, and simulation with conventional orthotropic material model, are compared. In simulation models, material properties were calibrated using uniaxial tension, bias extension and bending tests at 23° C., while the initial angle between the yarn direction and the global coordinates was defined as a material input property to identify the fiber layup. The simulation results in the upper-right quarter of FIG. 125 shows that the non-orthogonal material model established is capable of accurately predicting the physical experiments regarding the yarn angle distribution and blank draw-ins. For instance, the deviation of the maximum draw-in distance is about 9 mm, as listed in Table 9-1. For comparison, an orthotropic material model (MAT_002) is utilized in another simulation whose result is shown in the upper-left quarter of FIG. 189. Since the orthotropic model cannot track material property changes during the rotation of yarns, the corresponding simulation has a maximum draw-in deviation of 24 mm, as listed in Table 9-1, not capturing the overall process behavior.

In the non-orthogonal model, yarn angle is defined as an output variable, while MAT_002 does not have the capability for direct visualization. For clarity, Table 9-18 compares the resulting shear angles at various locations obtained from the experiment and simulations. Again, it shows that the current model has improved the prediction accuracy. The fiber orientation (yarn angle) prediction errors at only half of the locations reach the proposed 5%, which is unsatisfactory.

TABLE 9-18 Draw-in distance and yarn angle comparison between simulation and experiment: simulation results are from the experiment- based non-orthogonal material model and the conventional orthotropic material model (MAT_002). Comparison Draw-in A B C D E F Non-orthogonal 40.22 mm 89° 89° 71° 40° 45° 65° Orthotropic 25.00 mm 70° 85° 86° 47° 59° 77° Experiment 49.02 mm 80° 88° 71° 49° 56° 66°

The reason for the unsatisfactory fiber orientation prediction accuracy is that this non-orthogonal material model only utilizes experimental data from uniaxial tension (pure tension) and bias-extension (pure shear) tests. The coupling between tension and shear is neglected. For real prepregs, an increase of tension along the yarns will increase the contact force and friction force between warp and weft yarns, resulting in a shear resistance increase. This kind of mechanism is not simulated by the experiment-based non-orthogonal model. It is, however, captured by the multiscale preforming simulation method, where virtual material characterization is performed by experimentally calibrated mesoscopic RVE models that can be deformed to arbitrary strain. To demonstrate the improvement from multiscale modeling, its simulation result is compared against the one obtained from the previous non-orthogonal material model. Final prepreg geometry and yarn angle distribution results are demonstrated in panel (a) of FIG. 190 together with the real preformed part. The draw-in distance and the yarn angle at the sampling locations from the simulation and the experiment are listed in Table 9-19. The comparison indicates that this multiscale method with tension-shear coupling fulfils the proposed 5% error of fiber orientation (yarn angle) prediction at 5 out of 6 sampling points, and prediction errors at all sampling points are less than 8%.

TABLE 9-19 Draw-in distance and yarn angle comparison between the simulation and the experiment: The simulation results are from the new multiscale material model and the tension-shear decoupled material model. Comparison Draw-in A B C D E F Multiscale 42.25 mm 86° 88° 73° 54° 57° 67° Non-orthogonal 40.22 mm 89° 89° 71° 40° 45° 65° Experiment 49.02 mm 80° 88° 71° 49° 56° 66°

The punch force-displacement curves from the two simulation cases and the experiment are compared in panel (b) of FIG. 190. The plots demonstrate that the multiscale preforming simulation method predicts the punch force nearly the same as the experimental one compared to the simulation using the experiment-based non-orthogonal material model, which underestimates the experimental punch force by around 26%. The small discrepancy between the forces from the new simulation method and the experiment when the punch displacement reaches to over 70 mm may be caused by the negligence of the prepreg thickness variation by the shell elements in the simulation. The small force discrepancy when the punch displacement ranges from 20 to 50 mm, however, may result from the fact that the temperature at some locations of the prepreg has not completely reached 23° C. at the initial stage of the preforming, leading to softer material behavior compared to the one for the simulation. As a summary, this multiscale preforming simulation method with tension-shear coupling can predict the draw-in distance and the yarn angle variation on the preformed prepreg. More importantly, it also predicts the punch force history with high accuracy. This multiscale method, therefore, has stronger predictive capability than the experiment-based non-orthogonal mode does, and can serve as a powerful tool for part performance prediction, process parameters optimization, material design, and defect analysis for future preforming works.

After single layer preforming validation with a ±45° initial fiber orientation and 6 yarn angle measuring locations, further validation with different measuring locations, different initial fiber orientations, and different numbers of prepreg layers (0°/90° one layer, ±45° one layer, and 0°/90°/±45° two layers) were performed. Because of project time limitations, only the simulation results from the experiment-based non-orthogonal model are compared with the experimental ones. The multiscale simulation method was not applied to these configurations.

Simulation and experiment results with different initial fiber orientations are depicted in FIG. 191. The draw-in distance comparison is listed in Table 9-20. It can be seen that most of the draw-in distances predicted by simulation are within the ranges of the experiment results. The two largest deviations happened, however, in the x direction draw-in for −45/+45 initial fiber orientation in both single layer and double layers cases. It is also worth noting that for single layer −45/+45 setup, in experiments, specimen size in the x direction actually became larger (negative draw-in distance) after preforming, which is different from not only the simulation, but also common forming processes.

TABLE 9-20 Double-dome draw-in distance comparison. Initial exp × draw-in/ sim × draw-in/ exp × draw-in/ exp × draw-in/ orientation inch inch inch inch  0/90 0.70 about 1.34 1.50 1.27 about 1.71 1.73 −45/+45 −0.60 about −0.15 0.10 1.43 about 1.95 1.70 0/90 in 1.00 about 1.38 1.00 1.53 about 1.61 1.46 0/90/−45/+45 −45/+45 in 0.25 about 0.32 1.45 0.99 about 1.19 1.62 0/90/−45/+45

For warp and weft yarn angle distribution, results are shown in FIG. 192. It can be seen that for the single layer cases, the simulation agrees well with the experiment. At nearly all measuring locations simulation results are within experimental deviation and achieve the proposed 5% prediction error compared to the averaged experimental results. For the double layer, however, the discrepancy is larger. The most possible reason for this discrepancy is that the rapid drop of prepreg temperature, as shown in panel (b) of FIG. 190, results in resin melting and re-solidify between two prepreg layers, which causes very tacky or even disappeared prepreg interface, leading to much larger interaction strength and makes relative sliding between prepregs a lot more difficult compared to simulation. Actually, from an experimental aspect, the supplied prepreg in the invention has the best preforming temperature at 50 to 80° C. when the resin is molten but not cured. When the temperature drops down to around 23° C., the resin will be hard and very sticky, leading to undesired features such as edge breakage, discontinuous deformation, and out-of-plane deformation or even folding, as shown in FIG. 193. In the future, not only the temperature-dependent prepreg surface interaction should be taken into account in preforming simulation, but also it might be necessary to adjust the temperature of the tools by warming the coolant to the desired temperature. The purpose for this operation is not simply to reduce simulation and experiment discrepancies, but more importantly, it can improve the quality of the final parts.

The double-dome preforming benchmark test established can introduce complex double curvature features at the size of common automobile parts. Combined with the corresponding quantitative measurement of local prepreg temperature history, draw-in distance, local yarn angle, and forming force also developed in the invention, it serves as an effective experimental approach to validate the preforming simulation methods developed in the invention. Validation results indicate that the developed models can reach the proposed fiber orientation prediction error of less than 5% most of the time, guaranteeing the models' application potential.

Besides for calibrating the models in the invention, this double-dome benchmark test can also serve as validation for preforming simulation models developed in the future or from other researchers due to the fact that it considers most of the process parameters and provides the most important criteria for the final parts' performance. As a result, this approach enables researchers in both academic and industrial fields to test their preforming models in a reliable way, so it motivates the invention of accurate preforming simulation models that can help increase production and broaden application of advanced CFRPs, while benefiting environmental emission and fossil fuel control.

Moreover, the double-dome preforming tests performed in the invention provide important information about the production of high-quality parts. Temperature control in not only the heated oven but also with forming tools is essential for the resin to fully melt and cause small prepreg deformation resistance and small prepreg surface interaction, which are the keys to smooth and defect-free final parts.

Despite the fact that this exact setup of the double-dome preforming validation experiment is difficult to be commercialized directly, it facilitates commercialization of the preforming models developed in the invention for validation of these models' prediction accuracy. The quantitative measurement approaches for local temperature, draw-in distance, local yarn angle, and forming force can be transferred to other research teams in an open source form to establish a widely-accepted preforming simulation validation standard, which can accelerate development of other high-accuracy preforming models. Temperature control and monitor experience gained from this preforming experiment can be implemented into preforming presses for real production in the form of heated coolant and embedded thermocouples, to produce smooth and defect-free CFRP parts. In the part-level preforming model validation according to the invention, we established a double-dome benchmark test and corresponding quantitative measurement approaches and criteria. The benchmark tests validate the prediction capabilities of the preforming models we developed. Moreover, it provides a trustworthy approach to test models from other researchers, and gives insight guidance for the design of preforming facilities.

For experiments with multiple parameters that need to be considered, like this preforming benchmark test, it is essential to properly perform the design of the experiment to clearly study the effects from parameters, while keeping material and time consumption low. For preforming using prepregs, it is important to control not only initial prepreg temperature, but also tool temperature to ensure the resin keeps melting during the whole preforming processes for smooth and defect-free parts production.

The foregoing description of the exemplary embodiments of the present invention has been presented only for the purposes of illustration and description and is not intended to be exhaustive or to limit the invention to the precise forms disclosed. Many modifications and variations are possible in light of the above teaching.

The embodiments were chosen and described in order to explain the principles of the invention and their practical application so as to activate others skilled in the art to utilize the invention and various embodiments and with various modifications as are suited to the particular use contemplated. Alternative embodiments will become apparent to those skilled in the art to which the present invention pertains without departing from its spirit and scope. Accordingly, the scope of the present invention is defined by the appended claims rather than the foregoing description and the exemplary embodiments described therein.

Claims

1. A method for design optimization and/or performance prediction of a material system, comprising:

generating a representation of the material system at a number of scales, wherein the representation at a scale comprises microstructure volume elements (MVE) that are of building blocks of the material system at said scale;
collecting data of response fields of the MVE computed from a material model of the material system over predefined sets of material properties/constituents and boundary conditions;
applying machine learning to the collected data of response fields to generate clusters that minimize a distance between points in a nominal response space within each cluster;
computing an interaction tensor of interactions of each cluster with each of the other clusters;
manipulanting the governing partial differential equation (PDE) using Green's function to form a generalized Lippmann-Schwinger integral equation; and
solving the integral equation using the generated clusters and the computed interactions to result in a response prediction that is usable for the design optimization and/or performance prediction of the material system.

2. The method of claim 1, further comprising passing the resulted response prediction to a next coarser scale as an overall response of that building block, and iterating the process until a final scale is reached.

3. The method of claim 1, wherein the building blocks are defined by material properties and structural descriptors obtained by modeling or experimental observations and encoded in a domain decomposition of structures for identifying locations and properties of each phase within the building blocks.

4. The method of claim 3, wherein the structural descriptors comprise characteristic length and geometry.

5. The method of claim 1, wherein the boundary conditions are chosen to satisfy the Hill-Mandel condition.

6. The method of claim 1, wherein the collected data of response fields comprise a strain concentration tensor, a deformation concentration tensor, stress tensor including PK-I stress and/or Cauchy stress tensors, plastic strain tensor, thermal gradient, or the like.

7. The method of claim 1, wherein the machine learning comprises unsupervised machine learning and/or supervised machine learning.

8. The method of claim 1, wherein the machine learning is performed with a self-organizing mapping (SOM) method, a k-means clustering method, or the like.

9. The method of claim 1, wherein the clusters are generated by marking all material points that have the same response field within the representation of the material system with a unique ID and grouping material points with the same ID.

10. The method of claim 9, wherein the generated clusters is a reduced representation of the material system, which reduces the number of degrees of freedom required to represent the material system.

11. The method of claim 10, wherein the generated clusters are a reduced order MVE of the material system.

12. The method of claim 1, wherein the computed interaction tensor is for all pairs of the clusters.

13. The method of claim 1, wherein said computing the interaction tensor is performed with fast Fourier transform (FFT), numerical integration, or finite element method (FEM).

14. The method of claim 1, wherein the PDE is reformulated as a Lippmann-Schwinger (LS) equation.

15. The method of claim 14, wherein said solving the PDE with the LS equation is performed with arbitrary boundary conditions and material properties.

16. The method of claim 1, wherein the collected data of response fields, the generated clusters, and/or the computed interaction tensor are saved in one or more material system databases.

17. The method of claim 16, wherein said solving the PDE with the LS equation is performed in real-time by accessing the one or more material system databases for the generated clusters and the computed interaction tensors.

18. A method for design optimization and/or performance prediction of a material system, comprising:

performing an offline data compression, wherein original microstructure volume elements (MVE) of building blocks of the material system are compressed into clusters, and an interaction tensor of interactions of each cluster with each of the other clusters is computed; and
manipulanting the governing partial differential equation (PDE) using Green's function to form a generalized Lippmann-Schwinger integral equation; and
solving the integral equation using the generated clusters and the computed interactions to result in a response prediction that is usable for the design optimization and/or performance prediction of the material system.

19. The method of claim 18, further comprising passing the resulting response prediction to a next coarser scale as an overall response of that building block, and iterating the process until a final scale is reached.

20. The method of claim 18, wherein the building blocks are defined by material properties and structural descriptors obtained by modeling or experimental observations and encoded in a domain decomposition of structures for identifying locations and properties of each phase within the building blocks.

21. The method of claim 20, wherein the structural descriptors comprise characteristic length and geometry.

22. The method of claim 18, wherein the boundary conditions are chosen to satisfy the Hill-Mandel condition.

23. The method of claim 18, wherein said performing the offline data compression comprises:

collecting data of response fields of the MVE computed from a material model of the material system over a predefined set of material properties and boundary conditions;
applying machine learning to the collected data of response fields to generate clusters that minimize a distance between points in a nominal response space within each cluster; and
computing the interaction tensor is for all pairs of the clusters.

24. The method of claim 23, wherein the collected data of response fields comprise a strain concentration tensor, a deformation concentration tensor, stress tensor including PK-I stress and/or Cauchy stress tensors, plastic strain tensor, thermal gradient, or the like.

25. The method of claim 23, wherein the machine learning comprises unsupervised machine learning and/or supervised machine learning.

26. The method of claim 23, wherein the machine learning is performed with a self-organizing mapping (SOM) method, a k-means clustering method, or the like.

27. The method of claim 23, wherein the clusters are generated by marking all material points having the same response field within the representation of the material system with a unique ID and grouping material points with the same ID.

28. The method of claim 27, wherein the clusters is a reduced representation of the material system, which reduces the number of degrees of freedom required to represent the material system.

29. The method of claim 28, wherein the clusters are adapted as a reduced order MVE of the material system.

30. The method of claim 23, wherein said computing the interaction tensor is performed with fast Fourier transform (FFT), numerical integration, or finite element method (FEM).

31. The method of claim 23, wherein the PDE is reformulated as a Lippmann-Schwinger (LS) equation.

32. The method of claim 31, wherein said solving the PDE with the LS equation is performed with arbitrary boundary conditions and material properties.

33. The method of claim 23, wherein the collected data of response fields, the generated clusters, and/or the computed interaction tensor are saved in one or more material system databases.

34. The method of claim 33, wherein said solving the PDE with the LS equation is performed with online accessing the one or more material system databases for the generated clusters and the computed interactions.

35. A non-transitory tangible computer-readable medium storing instructions which, when executed by one or more processors, cause a system to perform a method for design optimization and/or performance prediction of a material system, wherein the method is in accordance with claim 1.

36. A computational system for design optimization and/or performance prediction of a material system, comprising

one or more computing devices comprising one or more processors; and
a non-transitory tangible computer-readable medium storing instructions which, when executed by the one or more processors, cause the one or more computing devices to perform a method for design optimization and/or performance prediction of a material system, wherein the method is in accordance with claim 1.

37. A material system database usable for conducting efficient and accurate multiscale modeling of a material system, comprising:

clusters for a plurality of material systems, each of which groups all material points having a same response field within a region within a microstructural volume element (MVE) of a respective material system with a unique ID;
interaction tensors, each of which represents interactions of all pairs of the clusters (regions with unique ID) for the respective material system; and
response predictions computed based on the clusters and the interaction tensors.

38. The material system database of claim 37, wherein the clusters are generated by applying machine learning to data of response fields of the MVE computed from a material model of the respective material system over a predefined set of material properties and boundary conditions.

39. The material system database of claim 38, wherein the interaction tensors are computed with fast Fourier transform (FFT), numerical integration, or finite element method (FEM).

40. The material system database of claim 39, wherein the responses predictions are obtained by solving a governing partial differential equation (PDE) with the LS equation using the clusters and the computed interactions.

41. The material system database of claim 40, wherein the responses predictions comprise at least stiffness, stress responses, damage initiation, fatigue indicating parameter (FIP), and/or thermal expansion.

42. The material system database of claim 37, being configured such that some of the response predictions are assigned as a training set for training a different machine learning that connects processes/structures to responses/properties of the material system directly without going through the clustering and interaction computing processes at all; and some or all of the remaining response predictions are assigned as a validation set for validating the efficiency and accuracy of the multiscale modeling of the material system.

43. The material system database of claim 37, being generated with predictive reduced order models.

44. The material system database of claim 42, wherein the predictive reduced order models comprise a self-consistent clustering analysis (SCA) model, a virtual clustering analysis (VCA) model, and/or an FEM clustering analysis (FCA) model.

45. The material system database of claim 37, being updatable, editable, accessible, and searchable.

46. A method of applying the material system database of claim 37 for design optimization and/or performance prediction of a material system, comprising:

training a neural network with data of the material system database; and
predicting real-time responses of the material system during a loading process performed using the trained neueral network, wherein the real-time responses are used for the design optimization and/or performance prediction of a material system.

47. The method of claim 46, further comprising performing a topology optimization to design a structure with microstructure information.

48. The method of claim 46, wherein the neural network comprises a feed forward neural network (FFNN) and/or a convolutional neural network (CNN).

49. A non-transitory tangible computer-readable medium storing instructions which, when executed by one or more processors, cause a system to perform a method for design optimization and/or performance prediction of a material system, wherein the method is in accordance with claim 46.

Patent History
Publication number: 20210357555
Type: Application
Filed: Sep 16, 2019
Publication Date: Nov 18, 2021
Inventors: Wing Kam LIU (Oak Brook, IL), Jiaying GAO (Evanston, IL), Cheng YU (Evanston, IL), Orion L. KAFKA (Enosburg, VT)
Application Number: 17/273,438
Classifications
International Classification: G06F 30/23 (20060101); G06F 30/27 (20060101);