Skip to main content
Erschienen in: Integrating Materials and Manufacturing Innovation 2/2019

Open Access 28.03.2019 | Technical Article

A Comparative Study of the Efficacy of Local/Global and Parametric/Nonparametric Machine Learning Methods for Establishing Structure–Property Linkages in High-Contrast 3D Elastic Composites

verfasst von: Patxi Fernandez-Zelaia, Yuksel C. Yabansu, Surya R. Kalidindi

Erschienen in: Integrating Materials and Manufacturing Innovation | Ausgabe 2/2019

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

search-config
loading …

Abstract

Reduced-order structure–property (S-P) linkages play a pivotal role in the tailored design of materials for advanced engineering components. There is a critical need to distill these from the simulation datasets aggregated using sophisticated, computationally expensive, physics-based numerical tools (e.g., finite element methods). The recent emergence of materials data science approaches has opened new avenues for addressing this challenge. In this paper, we critically compare the relative merits of the application of four distinct machine learning approaches for their efficacy in extracting microstructure-property linkages from the finite element simulation data aggregated on high-contrast elastic composites with different microstructures. The machine learning approaches selected for the study have included different combinations of local/global and parametric/nonparametric approaches. Furthermore, the nonparametric approaches selected for this study are based on Gaussian Process (GP) models that allow for a formal treatment of uncertainty quantification in the predicted values. The predictive performances of these different approaches have been compared against each other using rigorous cross-validation error metrics. Furthermore, their sensitivity to both the dataset size and dimensionality has been investigated.
Hinweise

Publisher’s Note

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

Introduction

Advances in materials science and engineering have served as critical enablers for technology leaps in multiple industries, including information technology, aerospace, automobile, and energy [14]. As specific examples, nickel-based superalloys have increased efficiency in gas turbines [3, 5, 6], amorphous soft magnetic materials have lowered transformation losses in the electrical power distribution systems [4], and Li-ion batteries have made possible zero-emission electric vehicles [7]. However, the successful deployment of advanced materials in commercial products following their initial discovery in the laboratory often takes one to two decades. In part, this is due to the lack of a rigorous mathematical framework for the design of new and improved materials customized for selected applications. This current deficiency has been recognized in multiple recent strategic initiatives aimed at realizing advanced materials with reduced cost and time [1, 8].
Rigorous approaches to materials design require computational strategies that efficiently solve inverse problems. More specifically, materials design aims to identify the combinations of material chemistries and manufacturing/process conditions that would result in a targeted combination of material properties in the final product. In most cases, the design space that needs to be explored to facilitate these inverse solutions spans an extremely large and high-dimensional space. Formally, in order to accomplish this grand challenge, one needs to first establish robust and reliable process-structure-property (PSP) linkages [914]. It is very important that these PSP linkages are established in forms that allow application of established design methodologies (i.e., inverse solutions) [1520].
In this work, we will focus our attention on the establishment of structure–property linkages. Generally speaking, process–structure linkages are significantly harder to establish as they demand attention to the details of material structure evolution during the imposed process conditions. We will also limit our attention here to the material structure at the mesoscale, generally referred to as the microstructure. Composite theories [2125] offer an important avenue to formulate computationally efficient microstructure-property linkages. However, their success has been largely limited to highly idealized microstructures that can be effectively quantified using a small number of parameters such as phase volume fractions and average shape factors. Most microstructures encountered in advanced materials defy such simplistic descriptions. Composite theories have also encountered significant challenges in arriving at sufficiently accurate predictions of the homogenized properties in high-contrast composites (these are composites whose microscale constituents exhibit widely varying local properties) [2629].
Finite element (FE) approaches offer a sophisticated toolset for incorporating the intricate details of 3D microstructures along with highly complex details of the governing physics in arriving at estimates of homogenized (macroscale) properties [21, 3033]. However, the use of this toolset demands significant computational resources. At the present time, these toolsets are not practically viable for addressing directly the material design problems demanding inverse solutions. Therefore, the most logical approach is to generate a sufficiently large collection of simulation datasets covering a wide variety of microstructures using the FE toolsets, and then extract computationally low-cost, reduced-order models from such datasets. Although this strategy requires modest computational resources in order to conduct the FE simulations on a variety of microstructures of interest, it incurs a one-time only computational effort. When properly designed and executed, this strategy has the potential to produce major time and cost savings in addressing the inverse problems central to the materials design efforts.
The approach described above based on reduced-order models has enjoyed remarkable success in low- to moderate-contrast composites [31, 3436]. The overall effort involved in establishing reduced-order microstructure–property linkages can be broken down conveniently into two main subtasks. The first subtask involves rigorous statistical quantification of the material microstructure (often referred to as feature engineering in data science terminology). Recent work [31, 35, 37, 38] has demonstrated that this subtask can be effectively handled through principal component analysis (PCA) of n-point spatial correlations computed for each microstructure included in the study. The second subtask involves model building using various machine learning approaches. For these models, the descriptors of the microstructure (e.g., the PC representations of their spatial correlations) serve as inputs and the effective macroscale properties of interest serve as outputs. The simplest of the model building approaches employs a large suite of easily accessible linear regression toolsets [31, 35, 39, 40] with or without regularization (especially needed when only a small number of data points are available). On the other hand, it is also possible to deploy highly sophisticated approaches based on convolution neural networks (CNN) [4144] that completely skip the feature engineering step (the first subtask mentioned above) and rely mainly on the power of a sophisticated network of models in capturing the intricate nonlinearities present in the model being built. However, CNN approaches require a fairly large number of data points before arriving at a reliable model with good predictive accuracy for new microstructures [41, 42].
The selection of the appropriate machine learning approach is often a nontrivial task, especially in addressing materials design problems. As already mentioned, the number of available data points for training and testing the model is one of the major considerations in this selection. Most regression techniques need a large number of training data points, and this requirement usually grows exponentially with the number of inputs (i.e., microstructure descriptors). Consequently, whenever possible, it is desirable to seek a low-dimensional representation of the inputs (e.g., the PC representations of spatial correlations used to represent the material microstructure). A second consideration in the selection of an appropriate machine learning approach is the availability of a suitable model form that could be justified by pre-existing domain knowledge (or the governing physics for the phenomena being studied). A third consideration in the selection of the appropriate machine learning approach lies in the need for rigorous quantification of the uncertainty in the prediction. In this regard, it should be noted that the error resulting from the use of the reduced-order model in place of the rigorous physics-based numerical simulation represents only one of the contributors to the uncertainty in the model predictions. This is because uncertainty in the prediction can also come from uncertainty in the governing physics or in the values of the parameters used in the physics-based model. For situations where it is important to quantify rigorously the uncertainty in the predictions, one generally prefers the use of approaches employing Bayesian inference [30, 36, 37, 45]. Based on these considerations, one typically selects a global/local (i.e., the predictions for new inputs are made either using all available data points or a selection of available data points in close proximity to the prediction data point), parametric/nonparametric approach (i.e., using a specified model form or not using any specified model form), or a Bayesian/non-Bayesian approach to model building. Given the complexities associated with each of the resulting choices, the selection of an appropriate machine learning toolset is often nontrivial in most applications. In the present application for extracting microstructure–property linkages in high-contrast elastic composites, there has not yet been a systematic study of the relative merits of the different approaches on the resulting models.
The difficulty in establishing reliable and robust structure–property linkages for high-contrast composites stems from the highly nonlinear and complex interactions that occur between the microscale constituents for any imposed macroscale loading condition. There currently do not exist any validated model forms that can accurately capture these complex interactions. Furthermore, the microstructure space (i.e., the complete set of all microstructures of interest) is an extremely large space. As a result, global approaches (a single model addressing the complete microstructure space) to model building have not yet yielded satisfactory results. Although CNN approaches [41, 42, 46] have shown significant promise, they place extremely high demands on computational resources. The computational burden involved in training a sophisticated CNN becomes even more dramatic if one considers exploring different combinations of layers used in CNN and retraining the network in the event of adding additional data points to the training ensemble.
The goal of this work is to compare objectively the efficacy of the different machine learning methods that adopt different combinations of local/global, parametric/nonparametric, and Bayesian/non-Bayesian approaches for establishing structure–property linkages. The remainder of this paper will be organized as follows. We first describe the dataset including an ensemble of microstructures with a wide range of morphological diversity and their analyses using previously established finite element methods. Then, the generation of microstructure features in the form of reduced-order representation of the 2-point spatial correlations of the microstructures in the ensemble studied here will be summarized. We will identify and review the specific modeling approaches that will be compared in this study. Finally, the relative benefits from the use of the different model building strategies for predicting the effective elastic properties of high-contrast composites will be critically assessed for accuracy and robustness.

Dataset Generation

The 3D material microstructure and its effective elastic stiffness predicted by the FE method constitute the input and output for the structure–property linkage to be built in this work. In other words, one microstructure and its FE-predicted effective stiffness constitute one data point for building and validating the desired reduced-order model. In order to facilitate this model building process, a large ensemble of voxelized 3D microstructures with broad morphological diversity were synthetically generated; these will be referred to as microscale volume elements (MVEs; [42, 46, 47]) in this paper. The workflow used to generate this ensemble of synthetic microstructures is depicted in Fig. 1. For simplicity, the workflow is illustrated with 2D microstructures. The process illustrated in this figure is trivially extended to the 3D microstructures generated for this study. The microstructure generation process starts by assigning a random number within the range of [0,1] to every voxel in the MVE (2D or 3D). Then, a Gaussian filter is applied on this random number field while treating boundaries as periodic to produce periodic (discretized) fields (i.e., the values wrap around the boundaries without any marked discontinuities). The desired discretized MVEs are obtained by simply thresholding the filtered image to attain a pre-selected volume fraction.
For the MVE generation process described above, the mean and covariance of the Gaussian filter determine the microstructure morphology obtained in the generated MVE. A 3D Gaussian filter with a mean μ (a 3X1 vector) and a covariance Σ (a 3 × 3 symmetric positive definite matrix) can be expressed mathematically as:
$$ h(\boldsymbol{x})=\frac{exp(-\frac{1}{2}(\boldsymbol{x}-\boldsymbol{\mu})^{T}\boldsymbol{\Sigma}^{-1}(\boldsymbol{x}-\boldsymbol{\mu}))}{\sqrt{\left( 2\pi \right)^{3}\left|\boldsymbol{\Sigma} \right|}} $$
(1)
The mean vector can be selected arbitrarily and is taken as a zero vector. The components of Σ play a critical role in controlling the shapes of the constituents (e.g., elliptical phase regions) in the microstructure. The off-diagonal components of Σ were assigned zero values for all microstructures generated in this study. This essentially implies that the principal frame of the microstructure is fully aligned with the sample axes of the MVEs. However, if one desires to produce microstructures whose principal axes are misaligned with the sample frame, it can be accomplished by using nonzero values for the off-diagonal terms in Σ. In the 2D example shown in Fig. 1, the Gaussian filter has a larger variance in the y direction. Hence, the filtered and the segmented images show an elongated feature in the y direction. Sixty-four different 3D Gaussian filters (using different values for the diagonal terms in Σ) were generated to include a wide range of microstructure morphologies in the generated MVEs. A total of 140 different 3D random number fields were generated and the 64 Gaussian filters were applied on each to provide a total of ≈ 8900 3D filtered images. Each filtered image (see top right plot of Fig. 1) was binarized using a threshold value designed to yield a selected volume fraction of the white phase sampled from a uniform distribution in the 25–75% range.
The material system of interest in this study is a high-contrast elastic 3D composite. The higher contrast in the individual properties of the microscale constituents in such composites is expected to lead to longer range complex and nonlinear interactions at the microscale. In prior works involving low- to moderate-contrast composites [48, 49], MVE sizes of 21 × 21 × 21 voxels were found to be adequate for capturing these microscale interactions accurately. For the high-contrast composites, significantly larger MVEs of size 51 × 51 × 51 voxels were employed. This will ensure that the MVEs can be treated as RVEs (representative volume elements) in the FE prediction of their effective stiffness. The microscale constituents (i.e., depicted as white and black phases in the segmented MVE) are both assumed to exhibit isotropic elastic response. A contrast of 50 was obtained by setting the Young’s moduli of white and black phases as E1 = 120 GPa and E2 = 2.4 GPa, respectively. On the other hand, Poisson ratios were kept the same for both phases, i.e., ν1 = ν2 = 0.3. The targeted property of interest for each MVE was selected as the effective stiffness parameter, C11,eff. The value of C11,eff for each MVE was obtained by executing a FE simulation imposing suitable periodic boundary conditions that would render all macroscale strain components except 〈ε11〉 0 (angular brackets represent the volume average). Other details of the FE simulations can be seen in prior reports [48, 49].
FE simulation of each MVE takes around 15 min with 4 CPUs and 32 GB of memory on a supercomputer. Since the number of distinct microstructural configurations is an extremely large set, it is not practical to explore the microstructure space with FE simulations. The only practical approach is therefore to first train a high-fidelity reduced-order model, and then to use that model to explore the large microstructure design space.

Microstructure Quantification Framework

We briefly present the details of the microstructure quantification framework employed in this work, which includes the use of the 2-point spatial correlations (often simply referred to as 2-point statistics) and PCA.

Two-Point Statistics

Two-point statistics capture the joint probability density of the occurrence of local states h and h (in our case, these reflect the white- and black-colored phases in the MVEs) at locations separated by a specified vector, \(\vec {r}\), in the microstructure. It is computationally efficient to define and use a discretized description of the two-point statistics as [35, 38, 41, 50, 51]:
$$ f_{\vec{r}}^{h h^{\prime}}=\frac{1}{\left|S_{\vec{r}}\right|}\sum\limits_{\boldsymbol{s}} m_{\boldsymbol{s}}^{h} m_{\boldsymbol{s}+{\vec{r}}}^{h^{\prime}} $$
(2)
where \({m_{s}^{h}}\) denotes a digital representation of the microstructure (reflects the volume fraction of the spatial voxel s occupied by the local state h) and \(\vec {r}\) represents a selected discretized vector. s should be treated as a three-dimensional index array for 3D microstructure volumes, i.e., s = {sx,sy,sz}; this allows us to conveniently identify the unique position of each voxel in the MVE. \(S_{\vec {r}}\) represents the total number of distinct valid trials that could be conducted in the selected microstructure by the different placements of the associated vector \(\vec {r}\) in the selected microstructure (the requirement for validity is that both the head and tail of the vector lie inside the given microstructure). For the periodic MVEs considered in this work, \(S_{\vec {r}}\) is equal to total number of voxels S in the MVE.
From Eq. 2, it is clear that only four sets of two-point statistics can be defined for a two-phase composite (since h and h can each take only two values corresponding to each of the two phases, respectively). Two of these correlations are called autocorrelations (when h = h), while the others are referred to as cross correlations. Furthermore, for a two-phase composite, only one autocorrelation needs to be computed as the other correlations can be trivially related to the selected autocorrelation [35, 52, 53]. An example MVE and the three orthogonal sections of its white–white autocorrelation are shown in Fig. 2. The center value in the autocorrelation corresponds to the zero vector (i.e., \(\vec {r} = 0\)) and reflects the phase volume fraction in the microstructures of the type employed in this study (in the MVEs generated for this study, each voxel is only allowed to be occupied by one distinct local state). Several other statistical measures of the microstructure are embedded in the two-point autocorrelation shown in Fig. 2 [38, 50, 54].
The expression in Eq. 2 is most efficiently computed by exploiting the properties of discrete Fourier transforms (DFTs) and the fast Fourier transform (FFT) algorithm [50, 52, 53, 55]. Calculation of the two-point statistics takes around 0.025 s with a single processor of 3.3 GHz speed and 4 GB memory on a standard desktop computer for each MVE used in this study.

Dimensionality Reduction

The two-point autocorrelation provides a rich representation of the spatial statistics in the MVE. In fact, the number of distinct statistics computed for the MVEs used in this study would be 51 × 51 × 51/2 ≈ 66326 dimensions (factor 2 is due to the symmetry in the two-point autocorrelation). Establishing reduced-order models with such a large set of features is a cumbersome task. In prior works [12, 13, 31, 34, 35, 38, 56], PCA has been successfully employed to reduce the number of features (i.e., feature engineering). PCA essentially rotates the coordinate system in which the microstructure statistics are defined. The new coordinate system is selected to maximize the capture of the variance in the dataset in the minimum number of terms. Therefore, one could argue that the data is now viewed from the most informative point of view. Moreover, the components of the new coordinate system are ordered to reflect the capture of the variance from the highest to the lowest components. In other words, the most significant feature in the ensemble is captured in the first principal component.
The two-point statistics of the microstructure indexed by (k) can be expressed in the principal component space as:
$$ f_{\vec{r}}^{(k)}\approx \sum\limits_{i = 1}^{R^{*}}\alpha_{i}^{(k)}\varphi_{i,\vec{r}}+\bar{f}_{\vec{r}} $$
(3)
where R reflects the number of principal components (PCs) retained in the reduced-order representation of the microstructure statistics. The value of R is usually very small compared to the number of two-point statistics computed for each microstructure in the ensemble studied. Equation 3 essentially represents an orthogonal decomposition of the original statistics, where αi denote the weights (also called PC scores) and \(\varphi _{i,\vec {r}}\) are the basis vectors. The PC scores for each microstructure \(\alpha _{i}^{(k)}\) can now serve as the objective low-dimensional representation of its two-point statistics. The value of R in Eq. 3 has usually only been in the range of 5–10 for most of the reduced-order PSP linkages built in prior work [12, 31, 35].
The principal component representation of the 8900 MVEs generated for this study is shown in Fig. 3 in the first three PCs, where each point (corresponding to each MVE) has been colored based on its FE-estimated effective stiffness. Note that visual inspection reveals that there is indeed a strong correlation between PC1 and the response C11,eff. This is because PC1 often carries the volume fraction information, which is known to have a strong influence on the effective stiffness of the MVE. However, it is also known that the volume fraction information alone is insufficient to produce sufficiently accurate predictions of the effective stiffness of new RVEs. Therefore, it is necessary to include additional PCs in building the desired structure–property linkage for the high-contrast composites.
For this study, R was systematically varied up to a maximum value of 60 to reflect the fact that the high-contrast composites are likely to need many more features in the establishment of accurate structure–property linkages. In other words, the first 60 PC scores of each MVE can potentially serve as the input for the reduced-order models built in this work.

Reduced-Order Model Building

In this work, we seek to study critically the efficacy of different combinations of local/global, parametric/ nonparametric, and Bayesian/non-Bayesian strategies in arriving at accurate and robust reduced-order structure–property linkages for two-phase high-contrast elastic composites. We will briefly review next the relevant model building approaches.

Parametric Regression

Regression is the statistical methodology associated with inference of a functional relationship between a response (i.e., output), y, and covariates (i.e., inputs), \(\boldsymbol {x} \in \mathcal {R}^{p}\) [57, 58]. For n observed response–covariate pairs, (x1,y1) ,..., (xn,yn), the response can be parametrically modeled as:
$$ y_{i} = g\left( \boldsymbol{x}_{i}; \boldsymbol{\beta} \right) + \epsilon_{i} i = 1,\ldots,n $$
(4)
where g is an assumed functional form, β are the corresponding regression coefficients, and 𝜖i is a random error term. The error term is usually associated with measurement errors present in physical experiments and are often taken to be independent and identically distributed (i.i.d.) normal random variables with zero-mean and σ2 variance (i.e., \(\mathcal {N}\left (0,\sigma ^{2}\right )\)). In linear regression, one minimizes the sum of the squares of the errors in the observations, which leads to closed-form solutions for β and σ2. Note that the parametric function implied in Eq. 4 can include nonlinear effects with the introduction of suitable regressors as:
$$ \begin{array}{@{}rcl@{}} g\left( \boldsymbol{x}_{i}; \boldsymbol{\beta} \right) &=& [1, \ldots,\psi^{d}(\boldsymbol{x})] \cdot [\beta^{0},\ldots,\beta^{d}]^{T}\\ &= &\boldsymbol{\Psi}^{T}(\boldsymbol{x}) \cdot \boldsymbol{\beta} \end{array} $$
(5)
The central challenge in many engineering problems is often in identifying the model function, g. For simple physical systems, the functional relationship is often established from the underlying physics. In the case of complex systems, the functional relationship is generally unknown. Furthermore, the response may be highly nonlinear involving potentially a large number of covariates. Another complication sometimes encountered in practice is associated with assumptions regarding the error term, 𝜖. Errors are often plagued by autocorrelations and may display heteroskedasticity, i.e., σ2 varies with x [59].
Within the framework of parametric regression, there are methods available to address the difficulties associated with identifying g. A naive approach is to a priori assume a rich set of candidate regressor functions, Ψ in Eq. 5, which can yield a sufficiently flexible model. However, if Ψ is too rich (i.e., d becomes too large), the regression task becomes prone to overfitting and the established model may therefore have poor predictive capabilities for new inputs. Regularization is often employed to alleviate this problem through the introduction of a penalty term to modify the least-squares objective function. Lasso (least absolute shrinkage and selection operator) and ridge regression (RR) are two of the popular regularization strategies [60] employed in most model building tasks. Lasso includes a penalty term proportional to the L1-norm of the regression coefficients ||β||1 [40]. The effect is that unimportant ψi’s are driven to βi = 0 yielding a lean model. Ridge regression instead utilizes a L2-norm which drives coefficients to be “small,” but not necessarily 0. A weighted combination of both norms, γ||β||1 + (1 − γ)||β||2, is known as elastic net (EN) [40].

Nonparametric Regression

Rather than focusing on identifying the explicit model form of g from the data, which enables future error-free predictions y(x) = g(x), nonparametric methods seek to predict y(x) directly from the observed data without necessitating estimations of the true model form of g [61]. Kernel regression is one such popular nonparametric technique where the predictor is of the form:
$$ y(\boldsymbol{x}) = \frac{{\sum}_{i = 1}^{n} K\left( \boldsymbol{x}_{i}-\boldsymbol{x};\boldsymbol{\theta}\right) y_{i}}{{\sum}_{i = 1}^{n} K\left( \boldsymbol{x}_{i}-\boldsymbol{x};\boldsymbol{\theta}\right)}. $$
(6)
Here, K is a kernel function and 𝜃 are the kernel’s associated hyperparameters [62]. Radial basis functions are a popular choice of kernel within which the isotropic Gaussian kernel is popularly used; \(K\left (\boldsymbol {x}_{i}-\boldsymbol {x};\boldsymbol {\theta }\right )=e^{-\theta \left ||\boldsymbol {x}_{i}-\boldsymbol {x}\right ||^{2}}\). Kernel methods obtain predictions by computing weighted means of the observed data where the weights are prescribed via the kernel. In the case of the Gaussian kernel, the hyperparameter, 𝜃, controls which observations are considered “close” to x via exponential decay. Hyperparameters are commonly selected via cross-validation. Gaussian process (GP) models are mathematically similar to kernel regression models with the exception that they are built on a statistical framework. This enables the definition of a likelihood function which may be optimized to estimate the unknown hyperparameters without the need for cross-validation.

Global vs Local Methods

Besides parametric and nonparametric approaches, another distinction exists in the model building efforts in the form of global and local strategies. The global approach seeks to identify a predictive model, via either parametric or nonparametric means, over the entire considered domain. Local approaches instead establish many different predictive models over the considered domain. The local approach tailors the model to the local data being considered. The result is a flexible model that can capture spatially varying correlations between observations. The local approach is analogous to methods in science/engineering which simplify the description of nonlinear functions/systems via local linearization (e.g., controls, optimization, numerical methods). Perhaps the most well known of the local approaches is LOESS or LOWESS (locally weighted scatterplot smoothing) [63, 64]. In this approach, the prediction at each x is obtained by considering a weighted regression problem where weights, wi, are 0 if xi are far from x and otherwise are appropriately weighed. Another ubiquitous local approach is the random forest model [65].

Gaussian Process Regression

Krige in 1951 published a statistical method for interpolating spatially distributed training data [66]. This method was further refined and formalized by Matheron in 1963, who referred to the model as kriging to honor Krige’s contribution[67]. In the late 1980s, these methods were introduced in the statistics community as Gaussian Process (GP) models, which were used to build surrogate interpolators for deterministic computer experiments [68]. GP models have recently also found a home in the machine learning community where they are utilized for both regression and classification [69].
In the regression setting, observations are assumed to follow the following statistical model:
$$ y_{i} = \mu + Z\left( \boldsymbol{x}_{i}; \boldsymbol{\theta} \right) + \epsilon_{i} i = 1,\ldots,n $$
(7)
where μ is the mean response, Z (xi;𝜃) is a weakly stationary GP with E[Z] = 0 and Var[Z] = τ2. To simplify notation Z (x;𝜃) will be denoted as Z (x) with the implied dependence on hyperparameters. As before, 𝜖i is assumed i.i.d. with E[𝜖i] = 0 and Var[𝜖i] = σ2. Unlike errors, realizations from the GP are not independent, i.e., Cov (Z(xi),Z(xj)) = τ2R(xixj) where R is the correlation function. The correlation function is a positive definite function which, like radial basis functions, specifies the decay in correlation as a function of relative distance. Note that this a priori assumption may be interpreted as Bayesian; rather than placing priors on model parameters, the prior is instead that the function itself is a realization from a GP. Although many correlation functions exist, perhaps the most common is the isotropic Gaussian function. The anisotropic Gaussian correlation function,
$$ R(\boldsymbol{h}) = exp\left( -\sum\limits_{i = 1}^{p} \theta_{i} {h_{i}^{2}}\right) $$
(8)
is often utilized to incorporate a separated correlation structure over each of the dimensions p. Therefore, the decay in correlation may vary differently in each dimension. In the GP modeling framework, 𝜃 are often referred to as the correlation length scales.
Consider that a noise-free prediction at y(x) is desired. Since all realizations from the GP are correlated the combined set of points (x1,y1) ,..., (xn,yn) , (x,y) are jointly distributed realizations from a n + 1 multivariate normal.
$$ \left( \begin{array}{c} y_{1} \\ {\vdots} \\ y_{n} \\ y \end{array} \right) = \mathcal{N} \left( \mu \boldsymbol{1}, \tau^{2} \left[\begin{array}{ll} \boldsymbol{R} & \boldsymbol{r} \\ \boldsymbol{r}^{T} & 1 \end{array}\right] + \sigma^{2} \left[\begin{array}{ll} \boldsymbol{I} & \boldsymbol{0} \\ \boldsymbol{0}^{T} & 0 \end{array}\right] \right) $$
(9)
where 1 is a vector of 1’s, 0 a vector of 0’s, pairwise correlation between data in the training set are captured by Rij = R(xixj), and correlations between the desired point and data points in the training set are captured by rT = [R(xx1),…,R(xxn)]T. Since (x1,y1) ,..., (xn,yn) have been observed, the distribution of Eq. 9 conditional on the observed data can readily be computed using the familiar conditional multivariate normal formula:
$$ \begin{array}{@{}rcl@{}} y(\boldsymbol{x})&=& \mathcal{N} \left( \hat{\mu},\hat{{\sigma}}^{2}_{y} \right)\\ \hat{\mu}&=&\mu + \boldsymbol{r}^{T} \left( \boldsymbol{R}+\frac{\sigma^{2}}{\tau^{2}} \boldsymbol{I}\right)^{-1}(\boldsymbol{y}-\mu\boldsymbol{1})\\ \hat{{\sigma}}^{2}_{y}&=&\tau^{2}\left[1-\boldsymbol{r}^{T}\left( \boldsymbol{R}+\frac{\sigma^{2}}{\tau^{2}} \boldsymbol{I}\right)^{-1}\boldsymbol{r}\right] \end{array} $$
(10)
where yT = [y1,y2,...,yn]. The result is that \(E[y(\boldsymbol {x})]=\hat {\mu }\) provides the estimated mean response and confidence in the estimate is represented by \(Var[y(\boldsymbol {x})]=\hat {\sigma }^{2}_{y}\). Models with a constant mean term are referred to as ordinary kriging. The quantity σ2/τ2 can be interpreted as the noise to signal ratio. If the noise dominates the underlying signal, σ2 > τ2, then \(\boldsymbol {R} + \frac {\sigma ^{2}}{\tau ^{2}} \boldsymbol {I}\) becomes increasingly diagonal which results in an inability to infer correlations between xi in the dataset. Note that in the statistics literature, particularly in the field of computer experiments, there exists a diverse set of more sophisticated GP models. Examples include composite models simultaneously employing two or more GPs [70], limit kriging with a nonconstant unknown mean [71], nonstationary covariance structures [72], and universal kriging where global mean behavior is captured through a generalized least-squares regression term [73]. Computational burden is incurred in computing \(\left (\boldsymbol {R}+\frac {\sigma ^{2}}{\tau ^{2}} \boldsymbol {I}\right )^{-1}\) but as the inverse is only a function of the already observed data, it may be pre-computed and stored. The mean term may be estimated from [69, 73]:
$$ \mu = \frac{\boldsymbol{y}^{T}\left( \boldsymbol{R}+\frac{\sigma^{2}}{\tau^{2}} \boldsymbol{I}\right)^{-1}\boldsymbol{1}}{\boldsymbol{1}^{T}\left( \boldsymbol{R}+\frac{\sigma^{2}}{\tau^{2}} \boldsymbol{I}\right)^{-1}\boldsymbol{1}} $$
(11)
The remaining unknown hyperparameters, {σ2,τ2,𝜃}, can be estimated using a maximum likelihood estimate (MLE) [69, 73]. Fully Bayesian approaches are also possible where hyperparameters are considered as random variables and therefore {σ2,τ2,𝜃} must be sampled from the corresponding hyperparameter posterior density. However, in practice, this complexity is often neglected in lieu of the simpler MLE approach. The optimization which yields estimates \(\left \{\hat {\sigma }^{2},\hat {\tau }^{2},\hat {\boldsymbol {\theta }}\right \}\) necessitates iteration which requires inversion of a n × n matrix; therefore, hyperparameter estimation can become computationally costly. For complex problems, a fully Bayesian approach may be computationally challenging since the unknown hyperparameters are jointly distributed and each Markov chain Monte Carlo (MCMC) sample would require the inversion of a large covariance matrix. In certain cases, the choice of appropriate hyperparameter priors enables the derivation of hyperparameter posterior marginal distributions, which facilitates efficient MCMC sampling [74]. Yet another strategy when considering large spatial datasets is to consider the use of reduced-order predictive processes which are built on a small set of “knots” rather than the full spatial dataset [75]. However, in practice, this complexity is often neglected in lieu of the simpler approach.
The utility of GP models in regression may be impaired when considering large datasets. Inversion of a dense n × n matrix incurs \(\mathcal {O}(n^{3})\) computational cost. Even for moderate-sized datasets (n > 1000), the cost becomes a limiting factor. Also, consider that if a separated correlation function is utilized then there are p unknown roughness parameters and the optimization routine must solve for p + 2 unknown hyperparameters. Even for moderate p, p > 10, this presents a challenge.
Rasmussen and Williams propose that large datasets can be accommodated through the use of low-rank approximations of \(\left (\boldsymbol {R}+\frac {\sigma ^{2}}{\tau ^{2}} \boldsymbol {I}\right )^{-1}\) [69]. The approximate inverse can therefore be utilized to estimate the global predictor y(x). Another approach follows the LOESS strategy where local predictors are generated for each desired x. Gramacy and Lee introduced treed GP models where the domain \(\mathcal {X} \in \mathcal {R}^{p}\) is tree partitioned and a separate GP predictor utilized in each leaf [76].
In the regression setting, parametric models quickly grow to become computationally unwieldy when considering many candidate functions in high-dimensional space; the total number of regression parameters will be \(1+p+\binom {p}{2}= 1+p+\frac {p(p-1)}{2}\). With the increasing model complexity, the problem quickly grows intractable. GP complexity, measured in terms of unknown hyperparameters, only grows linearly with p. Nevertheless, for large p, optimization may become increasingly difficult. One method to alleviate such issues is the use of additive GPs [77].

Local Gaussian Process Approximation

More recently, Gramacy and Apley proposed another local GP model where predictions are generated by building a GP around x according to a nearest neighbors (NN) or more optimal criteria [78]. GPs are well suited for regression as they offer sufficient flexibility for generalizing and inferring the underlying functional behavior while avoiding the curse of dimensionality. However, consider that the correlation length scales 𝜃 may vary in x (i.e. nonstationarity). Similarly, the variance terms may also vary in x (i.e. heteroskedasticity). The local approximation further bolsters the predictiveness of GPs by implicitly accounting for local variations in the structure of the data.
The locally approximated GP requires a design \(\mathcal {D}\) comprised of Nx points in the local domain of x to build a corresponding predictor. Nearest neighbors (NN) is a natural choice but is shown to be suboptimal [78]. Instead, the authors showed that an optimal design can be generated by considering an informative design criteria. One such criterion seeks to minimize the mean squared error (MSE) which is related to \(\hat {\sigma }_{y}^{2}\) in Eq. 10. A more computationally efficient criterion is considered which seeks to maximize information gained and is referred to as active learning Cohn (ALC) [78]. The design is generated via a greedy algorithm which begins with an initial design, \(\mathcal {D}_{o}\), of No NN points, and then sequentially adds the next optimal point subject to the selected criteria.

Results and Discussion

Four distinct reduced-order models were generated in this work for capturing the elusive linkages between the microstructure and the effective stiffness in high-contrast elastic composites. In building each model, both the dimensionality and the size of the dataset were systematically varied to compare and contrast the efficacy of each model building strategy. In the present work, the dimensionality is controlled by the number of PC weights (R) utilized. The size of the dataset is varied according to the fraction utilized (F). For instance, F = 0.2 implies that the reduced-order model is built using a randomly selected 20% fraction of the full dataset.
Model efficacy was quantified by considering a 10-fold cross-validation strategy. The employed validation strategy was established such that each of the reduced-order models was tested on unseen data over all N observations. In this way, we mitigate against certain validation realizations producing anomalous results due to chance predictions at “easy” or “hard” regions in \(\mathcal {X}\). The following strategy was employed: (1) the data is randomly partitioned into F and 1 − F fractions, (2) each of these fractions is divided into 10 folds, (3) the model is trained on 9 folds of the F fraction, (4) predictions are made on the left-out fold in F and one of the folds in the 1 − F fraction, and (5) the process is repeated until predictions are made at all N points. Note that some of the considered models require the estimation of hyperparameters using cross-validation (e.g., the regularization terms in the EN model). In these cases, the training folds of the F fraction are further “internally” split into cross-validation folds for parameter estimation.
The following four modeling strategies were selected for this study.
1.
A global parametric model was built using an elastic net (EN) regularization term. The two regularization hyperparameters were obtained using cross-validation. The candidate regressors included all polynomial terms up to and including 4th-order terms and all possible cross-terms.
 
2.
A local parametric model was built using a LOESS implementation. A tri-cube weight function was used to weigh the data. A linear model was used to parametrically describe local trends. Note that the original work was developed for one-dimensional x [63]. Adaptation of the technique to higher dimensions necessitates the definition of an appropriate distance measure. In this work, we use the mean Euclidean distance where the mean is taken over the number of dimensions considered (R).
 
3.
Gaussian process regression (GPR) was used to establish a nonparametric global model. The hyperparameters were inferred using an efficient subsampling procedure. M points were randomly sampled from the considered training data of size N. We used either M = int(N/10) or M = int(R× 10), whichever was smaller. The hyperparameters were estimated from this sample of size M and the process was repeated 20 times. The average of these quantities was used to build the GPR model. This simple procedure allows for efficient estimation of the correlation hyperparameters. Note that GPR predictions have a tendency to gravitate toward spatially close data. The GP kernel in Eq. 6 may be interpreted as a weighting function which assigns a weight of 1.0 to training points close to a desired prediction, and near-zero weight to training points far from a prediction. Predictions can be shown to consist of a weighted sum of training data in which weights favor nearby points. In this sense, GPR predictions can be thought of as “local” since they gravitate toward local spatial data. Our use of global refers to a single model description (single set of hyperparameters) that describes the data over the entire space considered.
 
4.
The local Gaussian process (laGP) regression model was built using using the laGP R-package [79]. We fixed the number of neighbors considered at 200. The active learning criterion (ALC) was used to select the “most informative” neighbors. As the distribution of our data (PC weights) in \(\mathcal {X}\) is unknown but is likely highly localized and nonlinear, we believe that the locally approximated GP model (laGP) is best suited for building an accurate predictor.
 
In this work, our choice of GP nonparametric regression models is motivated by our desire to not only make predictions but also derive interpretable physical knowledge. As will be shown, the values of the inferred hyperparameters will yield physical information about underlying spatial relationships. There are many other choices of nonparametric regression models, which we have considered in this work. Random forest models are ubiquitous [80], yet challenging to interpret. Bayesian adaptive regression trees are similar to random forest but include the use of a Bayesian regularization criterion, which bolsters predictiveness [81], but yet again is difficult to interpret. Therefore, although there are many classes of nonparametric regression models, we focus on GPs as they enable both future predictions and interpretations of existing data.
For this study, we selected the MSE and mean absolute error (MAE) as the error metric to evaluate predictive performance. MSE can be expressed as:
$$ MSE = \frac{1}{N}\sum\limits_{i = 1}^{N}\left( y(\boldsymbol{x}_{i})-\hat{y}(\boldsymbol{x}_{i})\right)^{2}, $$
(12)
where \(\hat {y}(\boldsymbol {x}_{i})\) is a prediction made on unseen data using the previously described cross-validation strategy. MAE is similar to Eq. 12; however, squared quantities are replaced by absolute values.
Table 1
Result summary
model
R
MAE (GPa)
MSPE (GPa)
EN
60
2.40
3.29
Linear-LOESS
60
4.05
5.29
GPR
60
2.25
3.12
laGP
15
1.87
2.68
3-D CNN
NA
1.04
NA
The MSE and MAE from the best of each of the models obtained in this study as well as results from a recent paper using 3D convolutional neural network (CNN)[42] are summarized in Table 1. The prior CNN work, however, only reported MAE values and therefore there is no MSE value for comparison. MSE is useful if one is interested in putting additional emphasis on the predictions with higher residuals. If there are a few predictions with significantly high residuals, MSE will quickly shift to larger values due to the squared residual term. MSE can be especially misleading in noisy observations. On the other hand, MAE weighs all predictions equally, meaning it does not penalize the predictions considered outliers as bad as MSE. MAE is easier to interpret since it provides the direct residuals while MSE gives a sense of spread of the predictions in terms of residuals and helps focus on outlier predictions. Note that the 3D CNN does not utilize 2-point statistics and instead directly considers the entire 51 × 51 × 51 voxel microstructure representation. This approach enables the possibility of inferring higher order statistics through the filter weights which the CNN estimates from the data [41]. Hence, the improvement in predictive accuracy obtained through the 3D CNN is evidence that perhaps important information is available in the higher order statistics. It should however be noted that deep learning approaches such as CNN require significantly larger amounts of training data compared to standard machine learning applications. Furthermore, the computational effort involved in CNN training is also significantly higher, and often requires GPU computing. This is because the CNN accepts the entire 3D microstructure as the input, not the low-dimensional measures of the microstructure as employed in this work. It is also important to recognize that the extracted CNN is not easily interpretable.
Note that there are varying degrees of correlation scale among the differing principal components as shown in Fig. 4. Consider that the median roughness parameter in the GPR model for PC1 is 𝜃1 ≈ 0.78. Therefore, for any two points in PC space, (x1,x2), the correlation function expressed as a product of exponentials is:
$$ \begin{array}{@{}rcl@{}} R\left( \boldsymbol{x}_{1},\boldsymbol{x}_{2}\right) &=& exp\left( -0.78 \left( x_{1,1}-x_{2,1}\right)^{2}\right) \cdot \\ &&\prod\limits_{i = 2}^{p}exp\left( -\theta_{i} \left( x_{1,i}-x_{2,i}\right)^{2}\right) \end{array} $$
(13)
The correlation between points differing by Δ1 = 0.78− 1/2 ≈ 1.13 in PC1 will be penalized multiplicatively by exp(− 1) ≈ 0.368. Consider now as an example the influence of PC8. Solving \(\theta _{8}{{\Delta }_{8}^{2}}= 1\) illustrates that to contribute the same amount of correlation decay \({\Delta }_{8} = \theta _{*}^{-1/2} = 3.1\). In other words, the data is much more sensitive to changes in PC1 than PC8. This agrees with the visually clear global trend in Fig. 3 where PC1 appears to have the greatest impact on the response variable when compared to PC2 and PC3. The significance of this observations is that predictions of y (x) will favor “borrowing“ information from points spatially close in PC1, PC2, PC3, PC6, etc. This inference indicates that those particular PC features contain important physical insight about the system. Analyzing the roughness parameters associated with each individual input variable for their relative importance to output is called automatic relevance determination (ARD) [57, 69]. ARD can be effectively used to eliminate the uncorrelated variables from the input space, which can help avoid overfitting.
The first, second, third, and sixth basis vectors are shown in Fig. 5. All basis vectors are scaled to the same color limits displayed by the color bar at the right side. The first basis vector exhibits an almost uniform distribution with a peak at the origin (corresponding to zero vector; this peak is too small to be seen in the figure). This is an indication that PC1 score correlates strongly with the volume fraction information. On the other hand, PC2 basis map has negative values in the short vector region indicating that it captures a decrease in the particle size. The PC3 basis map exhibits slightly positive values in the x direction, indicating an expansion of the particle shapes in this direction. The PC6 basis map exhibits high negative values in the x direction and high positive values in both y and z directions. This captures shape changes in particles which are being shrunk in the x direction and expanded in the yz plane. Since the effective property of interest is the normal component in x direction in the stiffness tensor (i.e., C11,eff), the morphological features identified in these PC scores seem to be the correct ones needed for building high-fidelity reduced-order models.
The efficiencies of different model building strategies explored in this study are summarized in Fig. 6. More quantitatively interpretable results are also shown in Fig. 7. The MSPE computed from the 10-fold cross-validation is shown as a function of the training fraction and number of considered principle components. One trend is clear across all considered models—there is a sharp increase in predictive accuracy when including PC6PC15 into the model. In Fig. 4, the correlation length scales suggest that the model is equally sensitive to PC6 (loading direction morphology) and PC1 (volume fraction). In addition, PC11PC13 also have some nonnegligible influence on the response. Therefore, the improvements obtained by including additional PC’s is not simply a matter of “including more data” but instead can be attributed to “including better physics.” Recall that PCA is an unsupervised method where the PC bases are established without considering the response variable. Therefore, it is possible that higher order PC’s, which are believed to be less important when establishing a PC representation, may actually be rather influential on the response.
Among the considered models, the best predictive performance was obtained by the laGP model. Furthermore, the laGP model is able to significantly outperform the other models at lower R. This indicates that the high-contrast elasticity dataset considered is both highly nonlinear and may be plagued by nonstationarity. A single structure–property reduced-order mapping over the entire domain \(\mathcal {X}\) is difficult to obtain and instead local nonparametric relationships better describe the linkage. Interestingly, the laGP model performs worse with increasing number of PC’s. In this case, we believe that this is due to overfitting. Recall that, for all laGP evaluations, the number of neighbors considered for establishing the local model was fixed to 200 points. For R = 60, this corresponds to about three data points per dimension. This is considered rather low and therefore we believe that prediction deteriorates with increasing R since the “true” hyperparameters cannot be inferred. However, scaling the number of neighbors with increasing R was observed to increase the computational burden.
Interestingly, the EN model appeared to perform well at a large R. Unfortunately, regression is iterative and requires inversion of an extremely large XTX matrix (X is the design matrix) whose size is determined by the number of features considered (columns in X). The number of features grows exponentially with increasing R, especially when considering cross-terms of polynomials up to the fourth order. Therefore, for larger values of R, the EN was not able to train in a timely manner. Compare this difficulty instead to the global GPR model which is bounded instead to the dataset size (N) and thus at worst requires inversion of a N × N matrix.
The LOESS model performs the worst from the considered models. The limitation in LOESS is twofold: (1) it does not utilize regularization and hence, to mitigate against overfitting, we limited the model to only consider linear terms; and (2) for high-dimensional data, it is nontrivial to compute distance-based weights. In the GP setting, distance is utilized to compute correlations, which can be thought of as weights, but there are correlation length-scale hyperparameters which allow for flexibility in establishing feature importance. In the standard LOESS implementation, there exists no analogue. For instance in LOESS, differences in PC1 are equally as impactful as differences in PC8. However, from Fig. 4, it is clear that PC1 is more “important.”

Conclusions

A number of parametric, nonparametric, local, and global models were considered for performing regression on a large, numerically generated, high-contrast, 3D elastic composite MVE dataset. Hyperparameters obtained from Gaussian Process regression aided in identifying important PC components that were otherwise deemed relatively unimportant by PCA. The PC basis for one such component corresponded intuitively with problem-specific structure–property-coupled physics. A locally approximate Gaussian Process model was found to yield the best predictive structure–property model. This local nonparametric model was found to significantly outperform other models particularly when considering more compact microstructure representations. This indicates that the nonparametric model is adept at exploiting information-rich features and that the high-contrast elasticity dataset likely exhibits nonstationary or heteroskedastic behavior which cannot be captured accurately using global approaches. The source of this anomalous behavior is likely associated with the highly complex response of the microstructures considered. Nonparametric methods were found to outperform parametric methods while avoiding difficulties associated with the curse of dimensionality.

Data Availability

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literatur
1.
Zurück zum Zitat Materials genome initiative for global competitiveness. National science and technology council executive office of the president (2011) Materials genome initiative for global competitiveness. National science and technology council executive office of the president (2011)
2.
Zurück zum Zitat van Schalkwijk W, Scrosati B (2007) Advances in lithium-ion batteries. Springer, Boston van Schalkwijk W, Scrosati B (2007) Advances in lithium-ion batteries. Springer, Boston
3.
Zurück zum Zitat Reed RC (2008) The superalloys: fundamentals and applications. Cambridge University Press, Cambridge Reed RC (2008) The superalloys: fundamentals and applications. Cambridge University Press, Cambridge
4.
Zurück zum Zitat Hasegawa R (2000) Present status of amorphous soft magnetic alloys. J Magn Magn Mater 215:240–245 Hasegawa R (2000) Present status of amorphous soft magnetic alloys. J Magn Magn Mater 215:240–245
5.
Zurück zum Zitat Furrer D, Fecht H (1999) Ni-based superalloys for turbine discs. JOM 51(1):14–17 Furrer D, Fecht H (1999) Ni-based superalloys for turbine discs. JOM 51(1):14–17
6.
Zurück zum Zitat Pollock T, Tin S (2006) Nickel-based superalloys for advanced turbine engines: chemistry, microstructure and properties. J Propul Power 22(2):361–374 Pollock T, Tin S (2006) Nickel-based superalloys for advanced turbine engines: chemistry, microstructure and properties. J Propul Power 22(2):361–374
7.
Zurück zum Zitat Lu L, Han X, Li J, Hua J, Ouyang M (2013) A review on the key issues for lithium-ion battery management in electric vehicles. J Power Sources 226:272–288 Lu L, Han X, Li J, Hua J, Ouyang M (2013) A review on the key issues for lithium-ion battery management in electric vehicles. J Power Sources 226:272–288
8.
Zurück zum Zitat Integrated computational materials engineering: a transformational discipline for improved competitiveness and national security. The National Academies Press, Washington (2008) Integrated computational materials engineering: a transformational discipline for improved competitiveness and national security. The National Academies Press, Washington (2008)
9.
Zurück zum Zitat Mueller T, Kusne AG, Ramprasad R (2015) Machine learning in materials science: recent progress and emerging applications. Rev Comput Chem 29:186–273 Mueller T, Kusne AG, Ramprasad R (2015) Machine learning in materials science: recent progress and emerging applications. Rev Comput Chem 29:186–273
10.
Zurück zum Zitat Yang N, Yee J, Zheng B, Gaiser K, Reynolds T, Clemon L, Lu W, Schoenung J, Lavernia E (2017) Process-structure-property relationships for 316l stainless steel fabricated by additive manufacturing and its implication for component engineering. J Therm Spray Tech 26(4):610–626 Yang N, Yee J, Zheng B, Gaiser K, Reynolds T, Clemon L, Lu W, Schoenung J, Lavernia E (2017) Process-structure-property relationships for 316l stainless steel fabricated by additive manufacturing and its implication for component engineering. J Therm Spray Tech 26(4):610–626
11.
Zurück zum Zitat Kalidindi S (2015) Hierarchical materials informatics: novel analytics for materials data. Elsevier, Boston Kalidindi S (2015) Hierarchical materials informatics: novel analytics for materials data. Elsevier, Boston
12.
Zurück zum Zitat Yabansu YC, Steinmetz P, Hötzer J, Kalidindi S, Nestler B (2017) Extraction of reduced-order process-structure linkages from phase-field simulations. Acta Mater 124:182–194 Yabansu YC, Steinmetz P, Hötzer J, Kalidindi S, Nestler B (2017) Extraction of reduced-order process-structure linkages from phase-field simulations. Acta Mater 124:182–194
13.
Zurück zum Zitat Gomberg JA, Medford AJ, Kalidindi S (2017) Extracting knowledge from molecular mechanics simulations of grain boundaries using machine learning. Acta Mater 133:100–108 Gomberg JA, Medford AJ, Kalidindi S (2017) Extracting knowledge from molecular mechanics simulations of grain boundaries using machine learning. Acta Mater 133:100–108
14.
Zurück zum Zitat Gorgannejad S, Gahrooei MR, Paynabar K, Neu R (2019) Quantitative prediction of the aged state of ni-base superalloys using pca and tensor regression. Acta Mater 165:259–269 Gorgannejad S, Gahrooei MR, Paynabar K, Neu R (2019) Quantitative prediction of the aged state of ni-base superalloys using pca and tensor regression. Acta Mater 165:259–269
15.
Zurück zum Zitat Olson GB (1997) Computational design of hierarchically structured materials. Science 277(5330):1237–1242 Olson GB (1997) Computational design of hierarchically structured materials. Science 277(5330):1237–1242
16.
Zurück zum Zitat Panchal JH, Kalidindi S, McDowell DL (2013) Key computational modeling issues in integrated computational materials engineering. Comput Aided Des 45(1):4–25 Panchal JH, Kalidindi S, McDowell DL (2013) Key computational modeling issues in integrated computational materials engineering. Comput Aided Des 45(1):4–25
17.
Zurück zum Zitat Nellippallil AB, Rangaraj V, Gautham B, Singh AK, Allen JK, Mistree F (2017) A goal-oriented, inverse decision-based design method to achieve the vertical and horizontal integration of models in a hot rod rolling process chain. In: ASME 2017 International design engineering technical conferences and computers and information in engineering conference. American Society of Mechanical Engineers, pp v02BT03a003–v02BT03a003 Nellippallil AB, Rangaraj V, Gautham B, Singh AK, Allen JK, Mistree F (2017) A goal-oriented, inverse decision-based design method to achieve the vertical and horizontal integration of models in a hot rod rolling process chain. In: ASME 2017 International design engineering technical conferences and computers and information in engineering conference. American Society of Mechanical Engineers, pp v02BT03a003–v02BT03a003
18.
Zurück zum Zitat Fullwood D, Niezgoda S, Adams BL, Kalidindi S (2010) Microstructure sensitive design for performance optimization. Prog Mater Sci 55(6):477–562 Fullwood D, Niezgoda S, Adams BL, Kalidindi S (2010) Microstructure sensitive design for performance optimization. Prog Mater Sci 55(6):477–562
19.
Zurück zum Zitat Sanchez-Lengeling B, Aspuru-Guzik A (2018) Inverse molecular design using machine learning: Generative models for matter engineering. Science 361(6400):360–365 Sanchez-Lengeling B, Aspuru-Guzik A (2018) Inverse molecular design using machine learning: Generative models for matter engineering. Science 361(6400):360–365
20.
Zurück zum Zitat McDowell DL, Kalidindi S (2016) The materials innovation ecosystem: a key enabler for the materials genome initiative. MRS Bulletin 41(4):326–337 McDowell DL, Kalidindi S (2016) The materials innovation ecosystem: a key enabler for the materials genome initiative. MRS Bulletin 41(4):326–337
21.
Zurück zum Zitat Mortazavi B, Baniassadi M, Bardon J, Ahzi S (2013) Modeling of two-phase random composite materials by finite element, mori–tanaka and strong contrast methods. Compos Part B Eng 45(1):1117–1125 Mortazavi B, Baniassadi M, Bardon J, Ahzi S (2013) Modeling of two-phase random composite materials by finite element, mori–tanaka and strong contrast methods. Compos Part B Eng 45(1):1117–1125
22.
Zurück zum Zitat Argatov II, Sabina FJ (2017) A two-phase self-consistent model for the grid indentation testing of composite materials. Int J Eng Sci 121:52–59 Argatov II, Sabina FJ (2017) A two-phase self-consistent model for the grid indentation testing of composite materials. Int J Eng Sci 121:52–59
23.
Zurück zum Zitat Duan H, Wang JX, Huang Z, Karihaloo BL (2005) Size-dependent effective elastic constants of solids containing nano-inhomogeneities with interface stress. J Mech Phys Solids 53(7):1574–1596 Duan H, Wang JX, Huang Z, Karihaloo BL (2005) Size-dependent effective elastic constants of solids containing nano-inhomogeneities with interface stress. J Mech Phys Solids 53(7):1574–1596
24.
Zurück zum Zitat Fu SY, Feng XQ, Lauke B, Mai YW (2008) Effects of particle size, particle/matrix interface adhesion and particle loading on mechanical properties of particulate–polymer composites. Compos Part B 39(6):933–961 Fu SY, Feng XQ, Lauke B, Mai YW (2008) Effects of particle size, particle/matrix interface adhesion and particle loading on mechanical properties of particulate–polymer composites. Compos Part B 39(6):933–961
25.
Zurück zum Zitat Trofimov A, Drach B, Sevostianov I (2017) Effective elastic properties of composites with particles of polyhedral shapes. Int J Solids Struct 120:157–170 Trofimov A, Drach B, Sevostianov I (2017) Effective elastic properties of composites with particles of polyhedral shapes. Int J Solids Struct 120:157–170
26.
Zurück zum Zitat Fullwood D, Adams BL, Kalidindi S (2008) A strong contrast homogenization formulation for multi-phase anisotropic materials. J Mech Phys Solids 56(6):2287–2297 Fullwood D, Adams BL, Kalidindi S (2008) A strong contrast homogenization formulation for multi-phase anisotropic materials. J Mech Phys Solids 56(6):2287–2297
27.
Zurück zum Zitat Pham D, Torquato S (2003) Strong-contrast expansions and approximations for the effective conductivity of isotropic multiphase composites. J Appl Phys 94(10):6591–6602 Pham D, Torquato S (2003) Strong-contrast expansions and approximations for the effective conductivity of isotropic multiphase composites. J Appl Phys 94(10):6591–6602
28.
Zurück zum Zitat Mikdam A, Makradi A, Ahzi S, Garmestani H, Li DS, Remond Y (2009) Effective conductivity in isotropic heterogeneous media using a strong-contrast statistical continuum theory. J Mech Phys Solids 57(1):76–86 Mikdam A, Makradi A, Ahzi S, Garmestani H, Li DS, Remond Y (2009) Effective conductivity in isotropic heterogeneous media using a strong-contrast statistical continuum theory. J Mech Phys Solids 57(1):76–86
29.
Zurück zum Zitat Adams BL, Kalidindi S, Fullwood D (2012) Microstructure sensitive design for performance optimization. Butterworth-Heinemann, Boston Adams BL, Kalidindi S, Fullwood D (2012) Microstructure sensitive design for performance optimization. Butterworth-Heinemann, Boston
30.
Zurück zum Zitat Fernandez-Zelaia P, Joseph VR, Kalidindi S, Melkote SN (2018) Estimating mechanical properties from spherical indentation using bayesian approaches. Mater Des 147:92–105 Fernandez-Zelaia P, Joseph VR, Kalidindi S, Melkote SN (2018) Estimating mechanical properties from spherical indentation using bayesian approaches. Mater Des 147:92–105
31.
Zurück zum Zitat Paulson NH, Priddy MW, McDowell DL, Kalidindi S (2017) Reduced-order structure-property linkages for polycrystalline microstructures based on 2-point statistics. Acta Mater 129:428–438 Paulson NH, Priddy MW, McDowell DL, Kalidindi S (2017) Reduced-order structure-property linkages for polycrystalline microstructures based on 2-point statistics. Acta Mater 129:428–438
32.
Zurück zum Zitat Li X, Xu Y, Chen S (2016) Computational homogenization of effective permeability in three-phase mesoscale concrete. Constr Build Mater 121:100–111 Li X, Xu Y, Chen S (2016) Computational homogenization of effective permeability in three-phase mesoscale concrete. Constr Build Mater 121:100–111
33.
Zurück zum Zitat Pinz M, Weber G, Lenthe W, Uchic M, Pollock T, Ghosh S (2018) Microstructure and property based statistically equivalent rves for intragranular γ- γ’microstructures of ni-based superalloys. Acta Mater 157:245–258 Pinz M, Weber G, Lenthe W, Uchic M, Pollock T, Ghosh S (2018) Microstructure and property based statistically equivalent rves for intragranular γ- γ’microstructures of ni-based superalloys. Acta Mater 157:245–258
34.
Zurück zum Zitat Latypov MI, Kalidindi S (2017) Data-driven reduced order models for effective yield strength and partitioning of strain in multiphase materials. J Comput Phys 346(13):242–261 Latypov MI, Kalidindi S (2017) Data-driven reduced order models for effective yield strength and partitioning of strain in multiphase materials. J Comput Phys 346(13):242–261
35.
Zurück zum Zitat Gupta A, Cecen A, Goyal S, Singh AK, Kalidindi S (2015) Structure–property linkages using a data science approach: application to a non-metallic inclusion/steel composite system. Acta Mater 91:239–254 Gupta A, Cecen A, Goyal S, Singh AK, Kalidindi S (2015) Structure–property linkages using a data science approach: application to a non-metallic inclusion/steel composite system. Acta Mater 91:239–254
36.
Zurück zum Zitat Jung J, Yoon JI, Park HK, Kim JY, Kim HS (2019) Bayesian approach in predicting mechanical properties of materials: Application to dual phase steels. Mater Sci Eng A 743:382–390 Jung J, Yoon JI, Park HK, Kim JY, Kim HS (2019) Bayesian approach in predicting mechanical properties of materials: Application to dual phase steels. Mater Sci Eng A 743:382–390
37.
Zurück zum Zitat Iskakov A, Yabansu YC, Rajagopalan S, Kapustina A, Kalidindi S (2018) Application of spherical indentation and the materials knowledge system framework to establishing microstructure-yield strength linkages from carbon steel scoops excised from high-temperature exposed components. Acta Mater 144:758–767 Iskakov A, Yabansu YC, Rajagopalan S, Kapustina A, Kalidindi S (2018) Application of spherical indentation and the materials knowledge system framework to establishing microstructure-yield strength linkages from carbon steel scoops excised from high-temperature exposed components. Acta Mater 144:758–767
38.
Zurück zum Zitat Altschuh P, Yabansu YC, Hötzer J, Selzer M, Nestler B, Kalidindi S (2017) Data science approaches for microstructure quantification and feature identification in porous membranes. J Membr Sci 540(1):88–97 Altschuh P, Yabansu YC, Hötzer J, Selzer M, Nestler B, Kalidindi S (2017) Data science approaches for microstructure quantification and feature identification in porous membranes. J Membr Sci 540(1):88–97
39.
Zurück zum Zitat Mangal A, Holm EA (2018) Applied machine learning to predict stress hotspots i: face centered cubic materials. Int J Plast 111:122–134 Mangal A, Holm EA (2018) Applied machine learning to predict stress hotspots i: face centered cubic materials. Int J Plast 111:122–134
40.
Zurück zum Zitat Friedman J, Hastie T, Tibshirani R (2001) The elements of statistical learning, vol 1. Springer series in statistics. New York, NY, USA Friedman J, Hastie T, Tibshirani R (2001) The elements of statistical learning, vol 1. Springer series in statistics. New York, NY, USA
41.
Zurück zum Zitat Cecen A, Dai H, Yabansu YC, Kalidindi S, Song L (2018) Material structure-property linkages using three-dimensional convolutional neural networks. Acta Mater 146:76–84 Cecen A, Dai H, Yabansu YC, Kalidindi S, Song L (2018) Material structure-property linkages using three-dimensional convolutional neural networks. Acta Mater 146:76–84
42.
Zurück zum Zitat Yang Z, Yabansu YC, Al-Bahrani R, Liao WK, Choudhary AN, Kalidindi SR, Agrawal A (2018) Deep learning approaches for mining structure-property linkages in high contrast composites from simulation datasets. Comput Mater Sci 151:278–287 Yang Z, Yabansu YC, Al-Bahrani R, Liao WK, Choudhary AN, Kalidindi SR, Agrawal A (2018) Deep learning approaches for mining structure-property linkages in high contrast composites from simulation datasets. Comput Mater Sci 151:278–287
43.
Zurück zum Zitat Ryczko K, Mills K, Luchak I, Homenick C, Tamblyn I (2018) Convolutional neural networks for atomistic systems. Comput Mater Sci 149:134–142 Ryczko K, Mills K, Luchak I, Homenick C, Tamblyn I (2018) Convolutional neural networks for atomistic systems. Comput Mater Sci 149:134–142
44.
Zurück zum Zitat Kondo R, Yamakawa S, Masuoka Y, Tajima S, Asahi R (2017) Microstructure recognition using convolutional neural networks for prediction of ionic conductivity in ceramics. Acta Mater 141:29–38 Kondo R, Yamakawa S, Masuoka Y, Tajima S, Asahi R (2017) Microstructure recognition using convolutional neural networks for prediction of ionic conductivity in ceramics. Acta Mater 141:29–38
45.
Zurück zum Zitat Fernandez-Zelaia P, Melkote SN (2019) Statistical calibration and uncertainty quantification of complex machining computer models. Int J Mach Tools Manuf 136:45–61 Fernandez-Zelaia P, Melkote SN (2019) Statistical calibration and uncertainty quantification of complex machining computer models. Int J Mach Tools Manuf 136:45–61
46.
Zurück zum Zitat Yang Z, Yabansu YC, Jha D, Liao WK, Choudhary AN, Kalidindi SR, Agrawal A (2019) Establishing structure-property localization linkages for elastic deformation of three-dimensional high contrast composites using deep learning approaches. Acta Mater 166:335–345 Yang Z, Yabansu YC, Jha D, Liao WK, Choudhary AN, Kalidindi SR, Agrawal A (2019) Establishing structure-property localization linkages for elastic deformation of three-dimensional high contrast composites using deep learning approaches. Acta Mater 166:335–345
47.
Zurück zum Zitat Yabansu YC, Kalidindi S (2015) Representation and calibration of elastic localization kernels for a broad class of cubic polycrystals. Acta Mater 94:26–35 Yabansu YC, Kalidindi S (2015) Representation and calibration of elastic localization kernels for a broad class of cubic polycrystals. Acta Mater 94:26–35
48.
Zurück zum Zitat Landi G, Niezgoda S, Kalidindi S (2010) Multi-scale modeling of elastic response of three-dimensional voxel-based microstructure datasets using novel dft-based knowledge systems. Acta Mater 58(7):2716–2725 Landi G, Niezgoda S, Kalidindi S (2010) Multi-scale modeling of elastic response of three-dimensional voxel-based microstructure datasets using novel dft-based knowledge systems. Acta Mater 58(7):2716–2725
49.
Zurück zum Zitat Kalidindi S, Niezgoda S, Landi G, Vachhani S, Fast T (2010) A novel framework for building materials knowledge systems. Comput Mater Continua 17(2):103–125 Kalidindi S, Niezgoda S, Landi G, Vachhani S, Fast T (2010) A novel framework for building materials knowledge systems. Comput Mater Continua 17(2):103–125
50.
Zurück zum Zitat Cecen A, Yabansu YC, Kalidindi S (2018) A new framework for rotationally invariant two-point spatial correlations in microstructure datasets. Acta Mater 158:53–64 Cecen A, Yabansu YC, Kalidindi S (2018) A new framework for rotationally invariant two-point spatial correlations in microstructure datasets. Acta Mater 158:53–64
51.
Zurück zum Zitat Niezgoda S, Kanjarla AK, Kalidindi S (2013) Novel microstructure quantification framework for databasing, visualization, and analysis of microstructure data. Integr Mater Manuf Innov 2(1):3 Niezgoda S, Kanjarla AK, Kalidindi S (2013) Novel microstructure quantification framework for databasing, visualization, and analysis of microstructure data. Integr Mater Manuf Innov 2(1):3
52.
Zurück zum Zitat Niezgoda S, Fullwood D, Kalidindi S (2008) Delineation of the space of 2-point correlations in a composite material system. Acta Mater 56(18):5285–5292 Niezgoda S, Fullwood D, Kalidindi S (2008) Delineation of the space of 2-point correlations in a composite material system. Acta Mater 56(18):5285–5292
53.
Zurück zum Zitat Fullwood D, Niezgoda S, Kalidindi S (2008) Microstructure reconstructions from 2-point statistics using phase-recovery algorithms. Acta Mater 56(5):942–948 Fullwood D, Niezgoda S, Kalidindi S (2008) Microstructure reconstructions from 2-point statistics using phase-recovery algorithms. Acta Mater 56(5):942–948
54.
Zurück zum Zitat Turner DM, Niezgoda S, Kalidindi S (2016) Efficient computation of the angularly resolved chord length distributions and lineal path functions in large microstructure datasets. Model Simul Mater Sci Eng 24(7):075,002 Turner DM, Niezgoda S, Kalidindi S (2016) Efficient computation of the angularly resolved chord length distributions and lineal path functions in large microstructure datasets. Model Simul Mater Sci Eng 24(7):075,002
55.
Zurück zum Zitat Brough DB, Wheeler D, Kalidindi S (2017) Materials knowledge systems in python—a data science framework for accelerated development of hierarchical materials. Integr Mater Manuf Innov 6(1):36–53 Brough DB, Wheeler D, Kalidindi S (2017) Materials knowledge systems in python—a data science framework for accelerated development of hierarchical materials. Integr Mater Manuf Innov 6(1):36–53
56.
Zurück zum Zitat Brough DB, Kannan A, Haaland B, Bucknall DG, Kalidindi S (2017) Extraction of process-structure evolution linkages from x-ray scattering measurements using dimensionality reduction and time series analysis. Integr Mater Manuf Innov 6(2):147–159 Brough DB, Kannan A, Haaland B, Bucknall DG, Kalidindi S (2017) Extraction of process-structure evolution linkages from x-ray scattering measurements using dimensionality reduction and time series analysis. Integr Mater Manuf Innov 6(2):147–159
57.
Zurück zum Zitat Bishop CM (2006) Pattern recognition and machine learning. Springer, Berlin Bishop CM (2006) Pattern recognition and machine learning. Springer, Berlin
58.
Zurück zum Zitat Witten IH, Frank E, Hall MA, Pal CJ (2016) Data mining: practical machine learning tools and techniques. Morgan Kaufmann, Cambridge Witten IH, Frank E, Hall MA, Pal CJ (2016) Data mining: practical machine learning tools and techniques. Morgan Kaufmann, Cambridge
59.
Zurück zum Zitat Binois M, Gramacy RB, Ludkovski M (2018) Practical heteroskedastic gaussian process modeling for large simulation experiments. J Comput Graph Stat 27(4):808–821 Binois M, Gramacy RB, Ludkovski M (2018) Practical heteroskedastic gaussian process modeling for large simulation experiments. J Comput Graph Stat 27(4):808–821
60.
Zurück zum Zitat Tibshirani R (1996) Regression shrinkage and selection via the lasso. J R Stat Soc Ser B Methodol 58 (1):267–288 Tibshirani R (1996) Regression shrinkage and selection via the lasso. J R Stat Soc Ser B Methodol 58 (1):267–288
61.
Zurück zum Zitat Härdle W (1990) Applied nonparametric regression 19. Cambridge University Press, Cambridge Härdle W (1990) Applied nonparametric regression 19. Cambridge University Press, Cambridge
62.
Zurück zum Zitat Kvam PH, Vidakovic B (2007) Nonparametric statistics with applications to science and engineering, vol 653. Wiley, New Jersey Kvam PH, Vidakovic B (2007) Nonparametric statistics with applications to science and engineering, vol 653. Wiley, New Jersey
63.
Zurück zum Zitat Cleveland WS (1979) Robust locally weighted regression and smoothing scatterplots. J Am Stat Assoc 74 (368):829–836 Cleveland WS (1979) Robust locally weighted regression and smoothing scatterplots. J Am Stat Assoc 74 (368):829–836
64.
Zurück zum Zitat Cleveland WS, Devlin SJ (1988) Locally weighted regression: an approach to regression analysis by local fitting. J Am Stat Assoc 83(403):596–610 Cleveland WS, Devlin SJ (1988) Locally weighted regression: an approach to regression analysis by local fitting. J Am Stat Assoc 83(403):596–610
65.
Zurück zum Zitat Ho TK (1995) Random decision forests. In: 1995 proceedings of the third international conference on document analysis and recognition, vol 1. IEEE, pp 278–282 Ho TK (1995) Random decision forests. In: 1995 proceedings of the third international conference on document analysis and recognition, vol 1. IEEE, pp 278–282
66.
Zurück zum Zitat Krige DG (1951) A statistical approach to some basic mine valuation problems on the witwatersrand. J South Afr Inst Min Metall 52(6):119–139 Krige DG (1951) A statistical approach to some basic mine valuation problems on the witwatersrand. J South Afr Inst Min Metall 52(6):119–139
67.
Zurück zum Zitat Matheron G (1963) Principles of geostatistics. Econ Geol 58(8):1246–1266 Matheron G (1963) Principles of geostatistics. Econ Geol 58(8):1246–1266
68.
Zurück zum Zitat Sacks J, Welch WJ, Mitchell TJ, Wynn HP (1989) Design and analysis of computer experiments. Stat Sci 4(4):409–435 Sacks J, Welch WJ, Mitchell TJ, Wynn HP (1989) Design and analysis of computer experiments. Stat Sci 4(4):409–435
69.
Zurück zum Zitat Rasmussen CE, Williams CK (2006) Gaussian processes for machine learning, vol 1. MIT Press, Cambridge Rasmussen CE, Williams CK (2006) Gaussian processes for machine learning, vol 1. MIT Press, Cambridge
70.
Zurück zum Zitat Ba S, Joseph VR et al (2012) Composite gaussian process models for emulating expensive functions. Ann Appl Stat 6(4):1838–1860 Ba S, Joseph VR et al (2012) Composite gaussian process models for emulating expensive functions. Ann Appl Stat 6(4):1838–1860
71.
Zurück zum Zitat Joseph VR (2006) Limit kriging. Technometrics 48(4):458–466 Joseph VR (2006) Limit kriging. Technometrics 48(4):458–466
72.
Zurück zum Zitat Tuo R, Wu CJ, Yu D (2014) Surrogate modeling of computer experiments with different mesh densities. Technometrics 56(3):372–380 Tuo R, Wu CJ, Yu D (2014) Surrogate modeling of computer experiments with different mesh densities. Technometrics 56(3):372–380
73.
Zurück zum Zitat Santner TJ, Williams BJ, Notz WI (2013) The design and analysis of computer experiments. Springer, New York Santner TJ, Williams BJ, Notz WI (2013) The design and analysis of computer experiments. Springer, New York
74.
Zurück zum Zitat Chen H, Loeppky JL, Welch WJ (2017) Flexible correlation structure for accurate prediction and uncertainty quantification in bayesian gaussian process emulation of a computer model. SIAM/ASA J Uncertain Quantif 5 (1):598–620 Chen H, Loeppky JL, Welch WJ (2017) Flexible correlation structure for accurate prediction and uncertainty quantification in bayesian gaussian process emulation of a computer model. SIAM/ASA J Uncertain Quantif 5 (1):598–620
75.
Zurück zum Zitat Banerjee S, Gelfand AE, Finley AO, Sang H (2008) Gaussian predictive process models for large spatial data sets. J R Stat Soc Ser B Stat Methodol 70(4):825–848 Banerjee S, Gelfand AE, Finley AO, Sang H (2008) Gaussian predictive process models for large spatial data sets. J R Stat Soc Ser B Stat Methodol 70(4):825–848
76.
Zurück zum Zitat Gramacy RB, Lee HKH (2008) Bayesian treed gaussian process models with an application to computer modeling. J Am Stat Assoc 103(483):1119–1130 Gramacy RB, Lee HKH (2008) Bayesian treed gaussian process models with an application to computer modeling. J Am Stat Assoc 103(483):1119–1130
77.
Zurück zum Zitat Duvenaud DK, Nickisch H, Rasmussen CE (2011) Additive gaussian processes. In: Advances in neural information processing systems, pp 226–234 Duvenaud DK, Nickisch H, Rasmussen CE (2011) Additive gaussian processes. In: Advances in neural information processing systems, pp 226–234
78.
Zurück zum Zitat Gramacy RB, Apley DW (2015) Local gaussian process approximation for large computer experiments. J Comput Graph Stat 24(2):561–578 Gramacy RB, Apley DW (2015) Local gaussian process approximation for large computer experiments. J Comput Graph Stat 24(2):561–578
79.
Zurück zum Zitat Gramacy RB (2015) lagp: large-scale spatial modeling via local approximate gaussian processes in r. Journal of Statistical Software (available as a vignette in the laGP package) Gramacy RB (2015) lagp: large-scale spatial modeling via local approximate gaussian processes in r. Journal of Statistical Software (available as a vignette in the laGP package)
80.
Zurück zum Zitat Breiman L (2001) Random forests. Mach Learn 45(1):5–32 Breiman L (2001) Random forests. Mach Learn 45(1):5–32
81.
Zurück zum Zitat Chipman HA, George EI, McCulloch RE, et al. (2010) BART: Bayesian additive regression trees. Ann Appl Stat 4(1):266– 298 Chipman HA, George EI, McCulloch RE, et al. (2010) BART: Bayesian additive regression trees. Ann Appl Stat 4(1):266– 298
Metadaten
Titel
A Comparative Study of the Efficacy of Local/Global and Parametric/Nonparametric Machine Learning Methods for Establishing Structure–Property Linkages in High-Contrast 3D Elastic Composites
verfasst von
Patxi Fernandez-Zelaia
Yuksel C. Yabansu
Surya R. Kalidindi
Publikationsdatum
28.03.2019
Verlag
Springer International Publishing
Erschienen in
Integrating Materials and Manufacturing Innovation / Ausgabe 2/2019
Print ISSN: 2193-9764
Elektronische ISSN: 2193-9772
DOI
https://doi.org/10.1007/s40192-019-00129-4

Weitere Artikel der Ausgabe 2/2019

Integrating Materials and Manufacturing Innovation 2/2019 Zur Ausgabe

Thematic Section: Additive Manufacturing Benchmarks 2018

Meshfree Simulations for Additive Manufacturing Process of Metals

Thematic Section: 5th World Congress on Integrated Computational Materials Engineering

Bayesian Framework for Inverse Inference in Manufacturing Process Chains

Thematic Section: 5th World Congress on Integrated Computational Materials Engineering

Cloud-Based Materials and Product Realization—Fostering ICME Via Industry 4.0

    Marktübersichten

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