You are on page 1of 43

Journal Pre-proof

Predicting effective thermal conductivity of fibrous and particulate composite materials


using convolutional neural network

Chengcheng Shen, Qiang Sheng, Haifeng Zhao

PII: S0167-6636(23)00250-8
DOI: https://doi.org/10.1016/j.mechmat.2023.104804
Reference: MECMAT 104804

To appear in: Mechanics of Materials

Received Date: 17 May 2023


Revised Date: 12 August 2023
Accepted Date: 9 September 2023

Please cite this article as: Shen, C., Sheng, Q., Zhao, H., Predicting effective thermal conductivity of
fibrous and particulate composite materials using convolutional neural network, Mechanics of Materials
(2023), doi: https://doi.org/10.1016/j.mechmat.2023.104804.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2023 Published by Elsevier Ltd.


1

1 Predicting Effective Thermal Conductivity of Fibrous and


2 Particulate Composite Materials using Convolutional Neural
3 Network

4 Chengcheng Shen1,2, Qiang Sheng2, Haifeng Zhao1,2,*


5 1
University of Chinese Academy of Sciences, Beijing, China
6 2
Key Laboratory of Space Utilization, Technology and Engineering Center for Space Utilization,
7 Chinese Academy of Sciences, Beijing, China
8 *Correspondence: hfzhao@csu.ac.cn

f
oo
9 Abstract

r
10
-p
Thermal insulation composite materials, such as aerogels and cellular foams, typically
re
11 exhibit great performance such as low thermal conductivity and complex microstructures at the
lP

12 nanometer or micrometer scale. However, traditional micromechanics-based methods like


na

13 homogenization and finite element analysis may not accurately predict their effective thermal
ur

14 conductivity due to a limited understanding of microscopic heat transfer mechanisms and the
Jo

15 vast amounts of multi-scale microstructural data. In this paper, the effective thermal conductivity

16 of composite materials is predicted using a convolutional neural network (CNN). To model the

17 microstructure of the composites, digital images are generated using the Quartet Structure

18 Generation Set (QSGS) method and the Random Generation-Growth Method (RGGM). The Lattice

19 Boltzmann Method (LBM) is then used to predict the effective thermal conductivities. The CNN

20 model is trained and validated using the microstructural images as input data and the conductivities

21 predicted by LBM as output data. Subsequently, Parametric studies are conducted to investigate the

22 material characteristics, including volume fraction and microstructural anisotropy. Additionally, the

23 CNN predictions are compared with the results of experiments, analytical models, and
1
2

1 computational methods. Finally, the proposed CNN model is utilized to predict the thermal

2 conductivity of materials with novel microstructures that are not included in the training set. By

3 capturing the scattering characteristics of heterogeneous materials through artificial intelligence, the

4 CNN model predicts thermal conductivity more efficiently than traditional methods. This highlights

5 the potential of machine learning to advance materials science and accelerate development of

6 materials with desired properties.

f
Keywords

oo
7

r
8 Thermal insulation composite materials, fibrous and particulate microstructure, Lattice
-p
9 Boltzmann method (LBM), convolutional neural network (CNN), effective thermal conductivity
re
lP

10 1 Introduction
na

11 Industries widely use nanoscale microstructured fibrous and particulate composite materials
ur

12 (Khan et al., 2018; Liu and Chen, 2014; Moore and Shi, 2014). Their outstanding performance,
Jo

13 including light weight, extremely low thermal conductivity, and superior mechanical properties, is

14 closely tied to their complex microstructure (Clyne and Hull, 2019; Ishizaki et al., 2013; Norouzi et

15 al., 2020). However, the complex microstructure of heterogeneous materials presents a challenge in

16 predicting their macroscopic properties. Thermal conductivity, a crucial property for thermal

17 insulating materials, depends on various controlling parameters, including the components,

18 microstructure, and interfaces. A thorough understanding of these factors is essential (He and Xie,

19 2015). This is crucial for the development of synthetic methods and industrial applications. However,

20 the heat transfer mechanism of materials with microstructures is highly complex, and the

2
3

1 distribution of the material's microstructure, the multi-scale nature of its phases, and the size effect

2 on heat transfer all need to be considered (Bi et al., 2014).

3 Researchers typically predict the effective thermal conductivity by analyzing physical models

4 or experimental data, such as effective medium theory (EMT) (Ma and Nan, 2014), homogenization

5 method (Zhao et al., 2006), Lattice Boltzmann Method (LBM) (M. Wang et al., 2007), hybrid

6 computing method (Demuth et al., 2014), or empirical models (Wang et al., 2018), etc. The EMT

7 and homogenization methods are cost-effective and computationally simple, however they only

f
oo
8 consider critical microstructural information, such as volume fraction or geometrical dispersion

r
9 -p
(Choy, 2015).The LBM method based on Boltzmann transport equation can capture the
re
10 characteristics of material microstructure (Kim et al., 2017; Zhou and Cheng, 2014), but the
lP

11 calculation cost is very high due to its iterative scheme (Xu et al., 2017). Experimental methods for

12 measuring thermal conductivity can be costly, especially when high-throughput analysis is required
na

13 (Ju et al., 2021).


ur

14 Recently, the use of machine learning (ML) methods has gained significant attention as a means
Jo

15 of establishing a numerical mapping between material characteristics at the microscale and physical

16 properties at the macroscale (Liu et al., 2021; Vasudevan et al., 2021). For example, Li et al. (Li et

17 al., 2020) utilized pattern recognition algorithms of digital images to extract geometric features of

18 different phases from scanning electron microscope (SEM) images of powder metallurgy nickel-based

19 superalloys, and constructed an Artificial Neural Network (ANN) to describe the relationship

20 between material hardness and dozens of microstructural features. Convolutional neural network

21 (CNN) method is superior to traditional methods in extracting microstructural characteristics, as it

22 has the ability to implicitly capture features that go beyond human prior knowledge (O’Shea and

3
4

1 Nash, 2015). Rabbani et al. (Rabbani and Babaei, 2019) presented a permeability calculation

2 workflow that couples pore network modeling with LBM, and in order to reduce the calculation

3 time of LBM, an ANN is trained to predict the permeability based on the rock cross–sectional image

4 features. Wei et al. (Wei et al., 2018) trained CNN models to predict the effective thermal

5 conductivity of porous composite materials, and their results showed that the CNN model was more

6 accurate and faster than EMT. Li et al. (Li et al., 2019) generated stochastic reconstructions of

7 mesoscale models from shale rocks and used finite element method (FEM) to compute effective

f
oo
8 moduli as training data for the CNN model. The trained CNN model demonstrated high accuracy

r
9 -p
and efficiency in predicting the mechanical properties of actual shale samples. For single-cell
re
10 mechanics, Nguyen et al. (Nguyen et al., 2023) used a CNN-based computational framework that
lP

11 can reproduce a physical microfluidic system to investigate the individual cell’s deformability, their

12 datasets were generated from high-fidelity fluid-structure interaction simulations as well as Lattice
na

13 Boltzmann method (LBM).


ur

14 In the literature above, the CNN models were only utilized to predict macroscopic properties
Jo

15 of materials with isotropy or inclusion of one single shape (Wang et al., 2020; Wei et al., 2018), or

16 the study of CNN models was limited to the heterogeneous microstructures with weak anisotropic

17 properties and the property reported in one single direction (Li et al., 2019; Kim et al., 2022). As a

18 result, the potential and scope of the CNN model may be limited for more generalized heterogeneous

19 materials.

20 In this study, a CNN model is proposed to establish a cross-scale relationship between the

21 microstructural characteristics and anisotropic thermal conductivities of heterogeneous materials

22 with hybrid types of inclusions (that is, more than one types of inclusions, or priori-trained

4
5

1 inclusions). Geometrical reconstruction methods, such as the Quartet structure generation set

2 (QSGS) method (M. Wang et al., 2007) and random generation-growth method (RGGM) (Xie and

3 He, 2016) are utilized to generate 2D microstructural images of fibrous and particulate composite

4 materials. The LBM is used to calculate the effective thermal conductivity of 2D materials in two

5 principal directions. Microscopic images are then used as inputs to train a CNN model, with the

6 corresponding thermal conductivities serving as outputs. After the training process is completed, the

7 CNN model can accurately predict the effective thermal conductivities in two directions for

f
oo
8 heterogeneous materials. Once the CNN model is ready, an investigation of the microscopic

r
9 -p
parameters that impact the thermal conductivity, such as material volume fraction and anisotropic
re
10 features, is performed. The results of the CNN model are then compared with those obtained from
lP

11 traditional models. Furthermore, both artificial materials with both fibrous and particulate

12 microstructures and real microstructures from SEM images of the material are analyzed using both
na

13 CNN and LBM. The error of CNN model is less than 6% and the predicting time is almost negligible.
ur

14 The CNN model can capture the microscopic characteristics of materials more accurately and
Jo

15 achieve high-fidelity predictions of thermal properties at a faster rate compared to traditional

16 theoretical or numerical calculation methods.

17 The structure of the paper is organized as follows: first, the data sets of microstructural images are

18 generated by QSGS and RGGM, and the corresponding effective thermal conductivities are calculated

19 by LBM. Then, the CNN model is established and trained, and the prediction and experimental validation

20 of real composite material are carried out. Next, the prediction results of the CNN model for samples

21 with varying microscopic parameters are compared with those obtained using the EMT or

22 homogenization method. At last, the trained CNN model is used to predict the thermal conductivity of

5
6

1 microstructural samples generated through numerical simulations or through image processing of SEM

2 images of real composite materials.

3 2 Methods

4 Most thermal insulation materials, such as fibrous and particulate composite materials, exhibit

5 specific microstructures, such as ordered pores, defects, inclusions, etc. For example, the

6 microstructure photos of polyimide fibrous aerogel (Qian et al., 2018) and silica aerogel

f
oo
7 (Mohammadian et al., 2018) are shown in Fig.1. These materials often show a variety of excellent

r
8 properties including extremely light density (Qian et al., 2018), ultra-low thermal conductivity (He
-p
9 et al., 2016), excellent acoustic absorption (Arenas and Crocker, 2010), etc. The microstructure of
re
particulate composite materials significantly influences their macroscopic properties. Thus, it is
lP

10

11 crucial to develop a parametric microstructure model for the materials of interest to comprehend the
na

12 relationship between the macroscopic structure and the microscopic properties of the material.
ur

13 In this paper, the numerical simulation method is employed to generate a microscopic model
Jo

14 that closely resembles the real material, and to calculate the effective thermal conductivity based on

15 physics. The simulation data is then used as training data for the CNN model to establish the

16 mapping relationship between the material's microscopic structure and its macroscopic thermal

17 conductivity.

6
7

(a) (b)
Figure 1. Illustration of typical thermal insulation materials, (a) polyimide fibrous aerogel (Qian
et al., 2018),(b) silica aerogel (Mohammadian et al., 2018)

f
oo
1 2.1 Microstructure datasets

r
2
-p
The numerical reconstruction methods QSGS and RGGM are used to generated microstructure
re
3 models, which are close to the real fibrous and particulate microstructures of thermal insulation
lP

4 composite materials. Then, the macroscopic effective thermal conductivity of each model is
na

5 calculated by LBM on the microstructure-based representative volume element (RVE). This


ur

6 prediction is considered to be as accurate as the experimental results (Chiavazzo and Asinari, 2010;
Jo

7 Fu et al., 2022; Jabbari et al., 2017; Zhai et al., 2018).

8 2.1.1 Parametric generation of microstructural images

9 Given the intricacy of real material microstructure, researchers have developed a variety of

10 methods to reconstruct microstructure through the analysis and statistical evaluation of materials

11 (Bostanabad et al., 2018; Shojaeefard et al., 2016). In this paper, QSGS method is used to simulate

12 the microstructure of particulate materials, and RGGM method is used to simulate the one of fibrous

13 materials. Both methods are employed to create tunable numerical models of complex

14 microstructures in the form of pixelated images by adjusting various parameters.

7
8

1 It is important to note that in this study, only two-phase composite materials are considered.

2 Regarding the 2D digital image as RVE of a specific particulate or fibrous material, each pixel

3 represents either a matrix or an inclusion element. Then, the data structure of a 2D image is a

4 mathematical matrix consisting of two values: either 0 or 1. The '0' phase represented is displayed

5 as black, while the '1' phase is white. The size of each image is set to 100 x 100 pixels.

6 For particulate composite material generated by QSGS, the core distribution probability 𝐶𝑑 is

7 used to control the density of the nucleus’s volume fraction, (that is actually area fraction in 2D).

f
oo
8 The real volume fraction of particle is 𝜙, and it is set to range from 0.1 to 0.8. The directional

r
9 -p
growth probability 𝐷𝑖 is to control the growth direction of the particle. Here 𝑖 = [0, 1, … , 8]
re
10 represents the direction of the 8 points around the center point, as shown in Fig.2(a), each small
lP

11 square represents a point. By adjusting the parameter 𝐷𝑖 , the distribution of particles can be

12 controlled to be either anisotropic or isotropic. In total, 1125 digital samples of particulate models
na

13 are generated by this method.


ur

14 For fibrous material generated by RGGM, the fiber volume ratio 𝜙 is set to range from 0.1 to
Jo

15 0.5. The dimensionless coefficient of fiber length 𝐿 is defined as the number of pixels of the actual

16 fiber length divided by the number of pixels of the model side length, it is set to range from 0.2 to

17 1. The fiber width 𝐷 is the number of pixels of the actual fiber width, it is range from 1 to 10. The

18 parameter 𝜃 is defined as the angle of axial direction of fiber and the direction of x-axis, as shown

19 in Fig.2(b), it is range from 0° to 90°. The 𝜃 of each fiber is varying between 𝜃1 and 𝜃2 ,

20 representing the maximum and minimum angel of the fiber direction respectively. By tuning the

21 parameters, 3531 samples of fibrous models are generated. In this paper, 4656 samples of particulate

22 and fibrous models are created in total.

8
9

f
oo
(a) (b)

Figure 2. Description of reconstruction methods, (a) QSGS method and (b) RGGM

r
method

1
-p
The representative images produced by both methods are shown in Fig.3. It can be seen that;
re
2 these methods are capable of generating materials with either isotropic or anisotropic types of
lP

3 microstructures and achieving specific macroscopic properties by adjusting model parameters. It is


na

4 evident from the results that these methods can create complex microstructures that mimic real
ur

5 materials. The corresponding parameters used to generate the images shown in Fig. 3 are listed in
Jo

6 Tables 1 & 2.

Figure.3. Representative images of reconstructed microstructure, (a) Images produced by


QSGS, (b) Images produced by RGGM

9
10

1
Figure
𝜙 𝐶𝑑 𝐷1,3 𝐷2,4 𝐷5,7 𝐷6,8
number
(a-1) 0.1076 0.001 0.3 0.3 0.3 0.3
(a-2) 0.2838 0.1 0.4 0.4 0.4 0.4
(a-3) 0.3049 0.003 0.01 0.01 0.4 0.01
(a-4) 0.4158 0.01 0.2 0.01 0.01 0.01
(a-5) 0.7064 0.01 0.4 0.4 0.4 0.4
2 Table 1. Model parameters in QSGS
3
Figure
𝜙 𝐿 𝐷 𝜃1 𝜃2
number
(b-1) 0.1000 0.2 2 30 90

f
(b-2) 0.2027 0.6 4 30 90

oo
(b-3) 0.3254 0.4 10 60 75
(b-4) 0.2506 0.7 2 0 50

r
(b-5) 0.4000 0.8 -p 1 0 90
4 Table 2. Model parameters in RGGM
re
5 2.1.2 Calculation of effective thermal conductivity by LBM
lP
na

6 To calculate the effective thermal conductivities of these material samples, only heat

7 conduction is considered, and heat radiation and convection are ignored. The heat transfer governing
ur

8 equations for two-phase composites without internal heat source are in the following forms
Jo

𝜕𝑇
(𝜌𝑐𝑝 )0 ( ) = 𝑘0 ∇2 𝑇 (1.)
𝜕𝑡
𝜕𝑇
(𝜌𝑐𝑝 )1 ( ) = 𝑘1 ∇2 𝑇 (2.)
𝜕𝑡

9 where subscript 0 denotes matrix phase and 1 denotes inclusion phase, which are corresponding

10 to the value of each pixel in the RVE. 𝜌 is the material density, 𝑐𝑝 is the specific heat capacity,

11 𝑘0 and 𝑘1 are the thermal conductivities, 𝑇 is the temperature, 𝑡 is time.

12 In this paper, the domain of all RVEs is square, and the boundary conditions are set as shown

13 in Fig.4(a). The effective thermal conductivity 𝑘𝑒 is calculated as follows

𝑘𝑒 = 𝑞𝐿/∆𝑇 (3.)

10
11

1 where 𝑞 is the steady heat, 𝐿 is the length of the model along the steady heat flux, ∆𝑇 is the

2 temperature difference between the left and right boundaries while the upper and lower boundaries

3 are set to adiabatic. The boundary conditions mentioned above are used to calculate the thermal

𝑦
4 conductivity in the x direction 𝑘𝑒𝑥 . When calculating the thermal conductivity in the y direction 𝑘𝑒 ,

5 the left and right boundaries are set to adiabatic, the upper and lower are prescribed with constant

6 temperatures.

f
r oo
-p
re
lP
na
ur
Jo

(a) (b)
Figure 4. (a) Schematic diagram of RVE with prescribed boundary conditions to calculate 𝑘𝑒𝑥 ,
(b) 2-D LBM thermal model with D2Q9 lattice

8 LBM is a mesoscopic approach based on the evolution of statistical distribution of lattices. It

9 is used to calculate the effective thermal conductivity at sub-continuum scale by solving Boltzmann

10 transport equation (Mohamad, 2019; J. Wang et al., 2007). In this paper, a two-dimensional nine-

11 speed D2Q9 LBM model is used, the lattice arrangements diagram is shown in Fig.4(b). The

12 mathematical matrix generated by QSGS and RGGM is transformed into the corresponding lattice

13 with element '0' representing matrix and '1' representing inclusions. The evolution equation is given

14 as

11
12

1 𝑒𝑞
𝑔𝛼 (𝐫 + 𝐞𝛼 𝛿𝑡, 𝑡 + 𝛿𝑡) − 𝑔𝛼 (𝐫, 𝑡) = − [𝑔𝛼 (𝐫, 𝑡) − 𝑔𝛼 (𝐫, 𝑡)] (4.)
𝜏

1 where 𝐫 is the location vector, 𝑡 is the real time, 𝛿𝑡 is the time step, 𝛼 is the lattice direction,

2 α = [0,1, … ,9] , 𝑔𝑒𝑞 is the equilibrium distribution of the evolution variable 𝑔𝛼 , 𝐞𝛼 is the

3 discrete velocity, as shown below

0 𝛼=0
𝑒𝑞
𝑔𝛼 ={ 1/6𝑇 𝛼 = 1, 2, 3, 4 (5.)
1/12𝑇 𝛼 = 5, 6, 7, 8
(0,0) 𝛼=0
𝐞𝛼 = { (cos 𝜃𝛼 , sin 𝜃𝛼 )𝑐, 𝜃𝛼 = (𝛼 − 1)𝜋/2 𝛼 = 1, 2, 3, 4 (6.)

f
√2(cos 𝜃𝛼 , sin 𝜃𝛼 )𝑐, 𝜃𝛼 = (𝛼 − 5)𝜋/2 + 𝜋/4 𝛼 = 5, 6, 7, 8

oo
4 𝜏 is the dimensionless relaxation time for each phase, which is determined by the material property,

r
5 as shown below -p
re
3 𝑘0
𝜏0 = + 0.5 (7.)
2 (𝜌𝑐𝑝 ) 𝑐 2 𝛿𝑡
0
lP

3 𝑘1
𝜏1 = + 0.5 (8.)
2 (𝜌𝑐𝑝 ) 𝑐 2 𝛿𝑡
na

6 where c is a pseudo sound speed, defined as δx/δt, and set to be 1. In this work, to insure 𝜏 to
ur

7 be within (0.5, 2) (J. Wang et al., 2007), (𝜌𝑐𝑝 )0 and (𝜌𝑐𝑝 )1 are set to 10000, the time step 𝛿𝑡 =
Jo

8 0.01 , the ratio of 𝑘0 and 𝑘1 is set to 1:10. Note that all thermal conductivities in this paper are

𝑦
9 dimensionless values. To achieve 𝑘𝑒𝑥 and 𝑘𝑒 of each RVE, it takes about 12 minutes to run one

10 LBM numerical scheme in a workstation with i7-11850H, 16 cores, 32G RAM, and an NVDIA

11 T1200 GPU.

12 After achieving the effective conductivities of all samples, the data distribution is plotted in

13 Fig.5. Each dot represents the effective thermal conductivity in both x and y directions for one RVE.

14 The bar charts at the top and right of Fig.5 represent frequency distribution histogram for effective

15 thermal conductivity of both directions, respectively. The effective thermal conductivity of the

12
13

1 samples closely follows the two-dimensional normal distribution, and the expected

2 dimensionless values in both directions are about 1.5. Especially, about 86% of the models had

3 effective thermal conductivity between 1 and 4. In Fig.5, the dots within the grey color band

4 represent RVEs with isotropic effective thermal conductivity. It means the difference of effective

5 thermal conductivities in x and y directions is within 5%. These samples are accounting for 20.12%

6 of all RVEs. In this work, all other RVEs are regarded as the anisotropic material.

f
r oo
-p
re
lP
na
ur
Jo

Figure 5. The distribution of effective thermal conductivities

7 2.2 CNN model

8 The traditional method of calculating composite materials provides a comprehensive

9 framework for studying the relationship between microstructure and physical properties. However,

10 this method is susceptible to the accuracy of microscopic modeling, the rationality of physical

11 equations, and the precision of the discretization technique. Additionally, it incurs high calculation

12 costs and low efficiency, and is not easily integrated with existing experimental or simulation data
13
14

1 for further prediction.

2 As a data-driven method, the ML approach can automatically establish computational models

3 based on training data, instead of relying on manual programs with prior knowledge. Therefore,

4 with the accumulation of experimental and simulation data for materials, ML can be employed to

5 accurately and quickly calculate the physical properties of composite materials, as well as to

6 establish correlations between microscopic and macroscopic properties of materials.

7 The Artificial Neural Network (ANN) has gradually evolved into an important branch of

f
oo
8 artificial intelligence algorithms. It is a typical supervised machine learning algorithm consisting of

r
9 -p
a large number of interconnected nodes (or neurons). The structure of one neuron is shown in
re
10 Fig.6(a). At each neuron, the calculation is performed as follows:
lP

𝐨j = 𝑓(∑) = 𝑓 [∑(𝒘𝑖𝑗 ∙ 𝐱 i + 𝐛j )] (9.)


𝑖=1
na

11 where 𝐱 i is the input, 𝐨j is the output, 𝑓 is the activation function, 𝒘𝑖𝑗 and 𝐛j represent
ur

12 weights and biases respectively.


Jo

13 The typical network structure of ANN is shown in Fig.6(b). ANN usually has three kinds of

14 layers: input layer, hidden layers and output layer. By training the neural network, the values of

15 weights and biases are constantly optimized to make the outputs of the network approximate to the

16 target values.

17 CNN is a type of neural network architecture frequently employed in deep learning. Similar to

18 ANN, CNN is composed of neurons that self-optimize through learning (O’Shea and Nash, 2015).

19 CNN is commonly utilized for image feature extraction and recognition due to two distinct types of

20 hidden layers not present in a typical ANN: convolutional layers and pooling layers.

21 An example of convolution kernel of 2x2 size is shown in Fig.6(c), the kernel has 8 weights
14
15

1 value and 1 bias value. By sliding the convolution kernel over the input image, the convolution

2 kernel elements are dotted with the input elements to obtain the output images. The size of the input

3 image is 5x5, and the size of the output image after the convolution process is 4x4. The structure of

4 pooling layer is similar to convolution layer, but it has no kernel, as shown in Fig.6(d). It is usually

5 calculated by simple operation such as maximum, minimum, or average.

6 Compared to fully connected neural network, the CNN model received the input information

7 of the image locally through convolution kernels. And the weights and biases are shared by kernels,

f
oo
8 so the network parameters are greatly reduced (Nielsen, 2015). These advantages make

r
9 -p
convolutional neural networks the most attractive deep learning neural network structures.
re
lP
na
ur
Jo

(a) (b)

(c) (d)
Figure 6. (a) Diagram of a single neuron, (b) typical ANN architecture, (c) convolution process,
(d) average pooling process

15
16

1 2.2.1 Model description

2 In this paper, to capture the geometric characteristics of microstructure, a CNN model is built

3 to calculate the effective thermal conductivity of reconstructed image models. The CNN model is

4 consisted of one input layer, 4 convolution layers, 4 average pooling layers, 2 full connect layers

5 and one output layer with 2 outputs, as shown in Fig.7. Relu function is selected as activation

6 function, as shown below

f
oo
𝑓(𝑧) = 𝑚𝑎𝑥(0, 𝑧) (10.)

7 The CNN model is based on the open source deep learning library TensorFlow for GPU

r
8
-p
(Michelucci, 2019) and Keras (Chollet, 2021; Gulli and Pal, 2017). The total trainable parameters
re
9 are about 3.5 × 105 , which is only about 35 times of one image pixels (100x100). All parameters
lP

10 in kernels and weights are initialized randomly.


na

11
ur
Jo

Figure 7. Schematic diagram of CNN model architecture

12 2.2.2 Evaluation criterions and training process

13 After the CNN model is initialized, the mean square error (MSE) is selected to be the loss

16
17

1 function in training. The MSE is defined as

1 𝑛
𝑀𝑆𝐸 = ∑ (𝐲i − 𝐲̂i )2 (11.)
𝑛 𝑖=1

2 where n is the amount of samples, 𝐲i represents the dimensionless values of effective thermal

3 conductivity calculated by LBM of each RVE, 𝐲̂i represents the dimensionless values of effective

4 thermal conductivity predicted by CNN model. In the training process, optimizer ‘Adam’ (Kingma

5 and Ba, 2014) is selected to tune parameters to minimize the MSE.

6 The training process is conducted on the workstation mentioned in section 2.1.2. The training

f
oo
7 time for a single CNN model is approximately 20 minutes. Once trained, the CNN model acquires

r
8 -p
optimal parameters, enabling it to accurately predict the effective thermal conductivity of
re
9 microstructure with minimal MSE, and the time to predict the effective thermal conductivity of one
lP

10 sample is negligible.

11 In the training process, the total 4656 RVEs are randomly divided into training and test set. The
na

12 number of samples in test set is fixed at 400, and the number of RVEs in different sizes of training
ur

13 sets varies from 1333, 1600, 2000, 2666 and 4000. After several simple trials, 100 epochs of training
Jo

14 are selected for different sizes of training sets, the MSE is shown in Fig.8(a). The MSE slightly

15 decrease with the increase of training set size. However, due to the limitation of data set acquisition

16 cost, it is impossible to increase the training set size without limitation. So the number of training

17 set is fixed at 4000 samples, and the number of test set is chosen at 400 for further model training.

18 In order to determine appropriate training epochs, the set training epochs are increased to 200.

19 The training loss curve is shown in the Fig.8(b). Obviously, the over fitting phenomenon occurred

20 when the training epochs are increased. So, 2 dropout layers are insert to the model, and the dropout

21 ratio is set to 0.2 (Srivastava et al., 2014). The training loss curve for model with dropout layers is

17
18

1 shown in the Fig.8(c). Comparing the two error curves, it can be seen that the CNN model with the

2 dropout layers alleviates the overfitting phenomenon and reduces MSE on the test set. And the early

3 stopping strategy (Prechelt, 2012) is used to further reduce the loss of test set. Finally, the best CNN

4 model with MSE of 0.365% is obtained.

f
r oo
-p
re
lP

(a)
na
ur
Jo

(b)

18
19

f
(c)

oo
Figure 8. (a) MSE between training set and CNN prediction with different size of training set,
(b) training loss without dropout layers, (c) training loss with dropout layers.

r
1 -p
The final CNN model is used to predict the 400 samples in test set, and the comparisons
re
2 between CNN results and LBM results are shown in Fig.9. The x-axis represents the value of the
lP

3 effective thermal conductivity predicted by LBM method, and y-axis represents the predicted value

4 of CNN model. The black line indicates the baseline when the predicted value from CNN model is
na

5 equal to that of LBM method.


ur

6 Besides MSE evaluation criterion, the regression coefficient R2 is also used to evaluate
Jo

7 prediction accuracy of CNN model. R2 is defined as

2
∑𝑛𝑖=1(𝐲i − 𝐲
̂i )2
𝑅 =1− (12.)
1
∑𝑛𝑖=1(𝐲i − ∑𝑛𝑖=1 𝐲i )2
𝑛

8 where n is the amount of RVEs, 𝐲i represents the dimensionless values of effective thermal

9 conductivity calculated by LBM of each sample, 𝐲̂i represents the dimensionless values of

10 effective thermal conductivity predicted by CNN model. The closer the value of the regression

11 coefficient R2 to 1 indicates that the CNN's prediction results are closer to the one of LBM.

12 As seen from Fig.9, the R2 values of the effective thermal conductivity are 0.9966 and 0.9965

13 in x and y direction, respectively, which are very close to 1. It means that this CNN model establishes
19
20

1 an accurate relationship of predicting the corresponding effective thermal conductivities of both x

2 and y directions from the microstructure of materials. Up to this point, the CNN model is proved to

3 be as accurate as LBM method in this paper, and the following chapters use this model for further

4 predictions.

f
r oo
-p
re
lP
na

(a)
ur
Jo

(b)
Figure 9. Comparisons between CNN and LBM predictions of test set

20
21

1 3 Results and discussion

2 Upon establishing the CNN model, it is utilized to predict the effective thermal conductivity

3 of diverse microstructure models. Firstly, a polymeric composite material containing particulate

4 inclusions is fabricated, and the CNN model's performance is first validated by real experimental

5 data. Then, the CNN is employed to predict the effective thermal conductivity of the parallel fiber

6 model and elliptic inclusion model and compared with the traditional method. Lastly, the CNN's

f
7 ability to predict the effective thermal conductivity of one artificial material with a hybrid fiber and

oo
8 particulate microstructure, as well as the real microstructure model, is investigated.

r
-p
3.1 Model validation with experimental data
re
9
lP

10 Before utilizing the CNN model for novel predictions, experimental data is generated involving
na

11 composite materials with specifically customized volume fractions of inclusions. This data serves

to validate both the LBM and CNN models. First, a PDMS(Polydimethylsiloxane) matrix composite
ur

12
Jo

13 is fabricated in the lab. This composite material contains inclusions of spherical and irregularly

14 shaped SiO2 particles with the volume fraction between 5-50%. The preparation process of the

15 composite material is referred to the experiment of Yang et al.(Yang et al., 2022). The SEM photos

16 of the materials are shown in the Fig.10(a). Utilizing SEM photos and raw material properties, the

17 QSGS method generates multiple RVEs that correspond to the material's microstructure, each

18 possessing distinct particle volume fractions. Subsequently, the composites' effective thermal

19 conductivities are assessed using the transient plane heat source method by the thermal analyzer

20 (Hot Disk TPS 2500S). To mitigate experimental discrepancies, three parallel tests are conducted

21 for each sample material. Analysis of both the material's SEM images and the results of the effective

21
22

1 thermal conductivity tests, conducted in various directions during the experiment, reveals the

2 isotropic nature of the prepared PDMS composite material. As a result, the macroscopic

3 conductivity of the material is investigated solely in one direction.

4 Next, using LBM to calculate the effective thermal conductivities of the PDMS composite

5 RVEs. The thermal conductivity of the PDMS matrix is 0.18 W/(m ∙ K) , and the one of SiO2

6 inclusions is 0.5 W/(m ∙ K). Totally, 4000 samples are used to train the CNN model for method

7 verification.

f
oo
8 The comparisons between real experimental data, CNN and LBM predictions of PDMS

r
9 -p
composite RVEs are shown in Fig.10(b). The prediction results of the CNN model are very
re
10 consistent with the LBM, and the R2 value is above 0.99. The maximum relative error between the
lP

11 CNN prediction results and the experimental test is 3.04%, and the average error is 1.41%. The

12 predictions of LBM and CNN model are both close to that of real experiment, which proves that the
na

13 data generated by LBM is reliable, and the CNN model can well capture the results of LBM.
ur
Jo

(a)

22
23

f
(b)

oo
Figure 10. (a) SEM images of PDMS composites with SiO2 particles volume fraction of 10%,
30%, 50%, (b) Comparisons between LBM, CNN predictions and experiment results of the

r
composites, error bars represent the standard deviation of three parallel tests.

1
-p
re
2 To understand the CNN model's ability to capture microstructure, partial feature maps of the
lP

3 convolutional layers of the CNN model are visualized, as shown in Fig.11. The low, medium and
na

4 high-level feature maps are from the first 3 convolutional layers, respectively. Although CNN's
ur

5 'understanding' of microstructure images is implicit, and the physical meaning corresponding to the
Jo

6 information contained in the feature maps is not yet clear (Zeiler and Fergus, 2014). The meaning

7 of the feature map can be speculated (Li et al., 2019), for example: the low-level feature map may

8 count the volume fraction of inclusions; the medium-level one may capture the edges of the

9 inclusions and determine the orientation of the microstructure; each pixel in the high-level feature

10 map may correspond to a small piece in the original image. After multiple convolutional and pooling

11 layers, the macroscopic properties of the material can be obtained from the information embedded

12 in the feature maps or weights and biases.

23
24

Figure 11. Visualization of feature map


1

2 3.2 Comparison of anisotropic materials with traditional

f
methods

oo
3

r
4 To further examine the ability of the CNN model to capture microstructure characteristics, the
-p
5 model's prediction outcomes are compared with those of an analytical or semi-empirical equation.
re
Two different types of anisotropic microstructures are considered in this section including fibrous
lP

7 materials with parallel fibers at different alignment angles and particulate materials with elliptical
na

8 inclusions (length-to-diameter ratio of 2) at different volume fraction. These samples are beyond
ur

9 training, validation, and test sets, and the effective thermal conductivities of samples are calculated
Jo

10 by CNN, analytical equations, and LBM. The error of prediction results of CNN or analytical

11 equations is defined as

|𝑘𝑒𝑃𝑟𝑒𝑑𝑖𝑐𝑡𝑒𝑑 − 𝑘𝑒𝐿𝐵𝑀 |
𝐸𝑟𝑟𝑜𝑟 = × 100% (13.)
𝑘𝑒𝐿𝐵𝑀

12 where 𝑘𝑒𝑃𝑟𝑒𝑑𝑖𝑐𝑡𝑒𝑑 represents the predicted effective thermal conductivity from CNN or analytical

13 equations, and 𝑘𝑒𝐿𝐵𝑀 represents the thermal conductivity calculated by LBM. Note that LBM

14 results are regard as accurate in this section.

24
25

1 3.2.1 Fibrous material with parallel fibers

2 In this case, the anisotropic fibrous microstructures with parallel fibers are considered. As

3 shown in Fig.12, ten RVEs of parallel fibrous material are generated. The parameters are set as

4 follows: the fiber volume ratio 𝜙 is 0.3, the dimensionless coefficient of fiber length 𝐿 is 0.4, the

5 fiber width 𝐷 is 1, and the angle 𝜃1 and 𝜃2 are set the same [0°, 10°, 20°, …, 90°]. The parameter

6 θ is defined as the angle between the fiber direction and the x-axis, the same as definition in section

f
oo
7 2.1.1. Here, the ratio of dimensionless values of thermal conductivities 𝑘𝑠 and 𝑘𝑔 is set to 10:1,

8 the same as section 2.1.2. In this case, only the effective thermal conductivity of x direction is

r
9
-p
predicted, because in θ between 0° and 90°, the thermal conductivity in the x and y directions is
re
10 corresponding.
lP
na
ur
Jo

Figure 12. Ten RVEs of fibrous material with parallel fibers

11

12 Jagjiwanram and Singh (Jagjiwanram and Singh, 2004) established one equation based on

13 EMT and heat transfer analysis from the series and parallel model, as follows

2 1⁄2
𝑘𝑒 = [(𝑘parallel ∙ cos 𝜃) + (𝑘series ∙ sin 𝜃)2 ] (14.)

14 Here 𝑘parallel and 𝑘series are the effective thermal conductivities that calculated by parallel model
25
26

1 and series model respectively, as shown in Eq. (15, 16). The parallel and series model provided

2 upper and lower limits, which are also named Voigt and Reuss bounds in the micromechanics

3 community (Hill, 1963).

𝑘𝑝𝑎𝑟𝑎𝑙𝑙𝑒𝑙 = 𝜙𝑘𝑠 + (1 − 𝜙)𝑘𝑔 (15.)

𝑘𝑠 𝑘𝑔
𝑘𝑠𝑒𝑟𝑖𝑒𝑠 = (16.)
(1 − 𝜙)𝑘𝑠 + 𝜙𝑘𝑔

4 Xie et al. (Xie and He, 2016) developed a numerical model by solving the radiative transfer

5 equation and finite volume method, for the combined conduction and radiation heat transfer of

f
oo
6 aerogel insulation composite material. According to their numerical model, an empirical formula is

r
7 -p
fitted to calculate the parallel distributed fibrous material, as expressed below:
re
𝑘𝑒 = (2.087 − 1.2544𝑒 −1/0.07646𝛾 )(1 − 𝑒 −𝐿/0.55 )(𝜆𝑘parallel
2 (17.)
−𝑘series )𝑒 −𝜃/(8.17+120.13/𝛾−115.32/𝛾 ) + 𝜆𝑘series
lP

8 where 𝛾 denotes the thermal conductivity ratio of fiber phase and matrix phase, 𝐿 is the fiber
na

9 length which is a relative ratio compared with material size.


ur

10 The variation trend of effective thermal conductivity with angle and the prediction results of
Jo

11 different models are shown in Fig.13(a).

12 Obviously, the effective thermal conductivity in the x direction 𝑘𝑒𝑥 decreases as the angle 𝜃

13 increases, but the shape of the curve is different. The prediction errors of models are shown in

14 Fig.13(b). It can be seen that CNN model exhibits better accuracy than Jagjiwanram’s or Xie’s

15 models. The average error of CNN model is the lowest, it is 1.57%, while 15.12% for

16 Jagjiwanram’s model and 20.25% for Xie’s model.

17 The thermal conductivity of the parallel fibrous material is influenced not only by the angle,

18 but also by microscopic characteristics such as the fiber aspect ratio. Jagjiwanram's model clearly

26
27

1 overlooks the microscopic features beyond the angle, resulting in inconsistency with real-world

2 scenarios. Xie's model's formula is built based on the outcomes of their numerical heat conduction

3 model, encompassing the ratio of fiber length to the overall RVE size. LBM utilizes the numerical

4 domain of a representative microstructure as inputs for computing the effective thermal

5 conductivity. It comprehensively captures the microstructural attributes to a greater extent than the

6 aforementioned two models.

f
r oo
-p
re
lP
na
ur

(a)
Jo

(b)
Figure 13. (a) Comparison of effective conductivity calculated by LBM, analytical models and
CNN, (b) Error of Jagjiwanram’s model, Xie’s model and CNN per Eq. (13).

27
28

1 3.2.2 Particulate material with aligned elliptical inclusions

2 In this section, particulate materials with aligned elliptical inclusions are generated by a new

3 method. This method randomly generates the seeding location of each elliptical inclusion and

4 subsequently determinates whether the points near the dot are within the ellipse, marking those

5 within the ellipse as ‘1’ and those outside as ‘0’. The volume fraction is counted by pixels. The

6 parameters in the code are as follows: 𝐴 is the pixel number of major axis of elliptical particles and

f
oo
7 set to 15 pixels; 𝑎 is defined as the aspect ratio of the minor axis of the particle to the major axis,

8 it is set to 0.33; volume fraction 𝜙 is ranging from about 0.05 to 0.4. Similar to QSGS method,

r
9
-p
whether particles overlap each other is not considered. the ratio of dimensionless values of thermal
re
10 conductivities 𝑘0 and 𝑘1 is set to 1:10. One example of generated RVE is shown in Fig.14(a).
lP

11 After model generation, CNN, LBM and the following micromechanics methods are used to predict
na

12 the effective thermal conductivities of x and y directions.


ur

13 For the elliptical inclusions have the same form with aspect ratio 𝑎 and they are aligned with
Jo

14 the x-axis. Hashin and Shtrikman (Hashin and Shtrikman, 1963) proposed elaborated bounds for a

15 two-phase isotropic composite. Hashin and Shtrikman (H-S) model is based on analytical

16 micromechanics and variational approach, as follows:

1 − 𝑘𝑠 + 𝜙𝑘𝑠 + 𝑎𝑘𝑠
𝑘𝑒𝑥 = (18.)
1 + 𝑎(𝜙 + 𝑘𝑠 − 𝜙𝑘𝑠 )
𝑦 𝑎(1 − 𝑘𝑠 + 𝜙𝑘𝑠 ) + 𝑘𝑠
𝑘𝑒 = (19.)
𝑎 + 𝜙 + 𝑘𝑠 − 𝜙𝑘𝑠

17 The effective thermal conductivity results predicted by CNN model and H-S model are shown

18 in Fig.14(b). It can be seen that the predictions of H-S model are more accurate for low volume

19 fraction than for high volume fraction. But the volume fraction has no effect on CNN model

20 accuracy, as shown in Fig.14(c), the CNN model keeps a low prediction error for both low and high
28
29

1 volume fraction. The average error of CNN model is 0.58% for x direction, 2.25% for y direction,

2 but the error of H-S model is 6.10% and 4.96%, respectively.

f
oo
(a)

r
-p
re
lP
na
ur
Jo

(b)

(c)
Figure.14 (a) RVE of particulate material with aligned elliptical inclusions, (b) comparation of

29
30

analytical models and CNN, (b) Error of H-S model and CNN per Eq. (13).

1 Furthermore, the H-S model lacks parameters that describes the size of each particle, such as

2 the length of the long axis of an ellipse. As a matter of fact, due to scale effects, particles of different

3 sizes can result in varying effective thermal conductivities, even if they possess the same volume

4 fraction.

5 3.3 Model extrapolation

f
6 Regarding the CNN model trained in this study, it does not differentiate between particulate

oo
7 and fibrous microstructures. The remarkable capability of the CNN model is its ability to thoroughly

r
8
-p
capture the microstructure of digital RVEs and map them onto the material's macroscopic properties.
re
9 3.3.1 Artificial material - hybrid fibrous and particulate models
lP
na

10 To further explore the potential of the CNN model, RVEs of hybrid fibrous and particulate

microstructures are generated, then the trained CNN model is employed to predict their effective
ur

11
Jo

12 conductivities. Those hybrid RVEs are obtained by combining the models generated the QSGS and

13 RGGM methods through matrix addition operations. For example, as shown in Fig.15(a), elements

14 in the matrix part of the particulate model are replaced with elements in the fibrous model at the

15 corresponding position. Fifty anisotropic and isotropic RVEs are generated, they contain four matrix

16 operation methods: (i)only retain the internal structure of the particles, (ii)retain the external

17 structure of the particles, (iii)completely merge the particles and fibers, and (iv) completely subtract

18 particles and fibers.

19 Some RVEs as shown in Fig.15(b). And the effective thermal conductivities are predicted by

20 both LBM and CNN. Fig.15(c) shows the comparison between CNN and LBM predictions. The R2

30
31

1 values of x and y directions are all over 0.99. The maximum absolute value of the error is 4.87%,

2 the average absolute error of CNN model is 1.38% for x direction, 1.62% for y direction.

(a)

f
oo
(b)

r
-p
re
lP
na
ur
Jo

(c)
Figure.15. (a) Hybrid RVE generated by the addition of QSGS and RGGM models, (b) Five
hybrid RVEs, (c) Comparison between CNN and LBM predictions.

3 3.3.2 SEM images from real materials

4 The second case to explore the CNN model is using SEM images of real materials to reflect

5 the true microscopic characteristics. Due to the limitation of SEM images acquisition, the study is

6 not limited to the thermal insulation composite materials. These images are selected from SEM

31
32

1 images of polyimide fibrous aerogel (Qian et al., 2018) and ceramic matrix fiber composites (Justin

2 et al., 2020; Larson and Zok, 2018). These SEM images are binarized (Otsu, 1979) to achieve

3 images in the grayscale. The image size is modified to 100x100 pixels and used as the input for

4 CNN model, as shown in Fig.16(a). Note that, for ceramic matrix composites, fibers are considered

5 as phase ‘1’ and shown as white points, the ceramic is phase ‘0’ and shown as black points.

6 The effective thermal conductivities in x and y directions are also predicted by LBM. Then the

7 predictions are compared between CNN and LBM, as shown in Fig.16(b). The maximum relative

f
oo
8 error of CNN model is 5.69%, the average relative error is 2.66% for x direction, 2.48% for y

r
9 -p
direction. Despite the complex microstructure of real materials, the CNN model is still able to make
re
10 accurate predictions. Its relative error is less than 6%.
lP

11
na
ur
Jo

32
33

(a)

f
oo
(b)

r
Figure.16. (a) SEM images of real materials and the workflow of image processing, (b)
-p
comparison between CNN and LBM predictions.
re
1 The CNN model effectively achieves precise predictions for both hybrid fibrous and particulate
lP

2 models, as well as real material microstructures derived from SEM images. However, it is crucial

3 to reiterate that the training data for this CNN model encompasses solely two specific types of
na

4 microstructure inclusions and lacks representation of hybrid training samples. This underscores the
ur

5 distinctive "comprehension capabilities" exhibited by machine learning models compared to those


Jo

6 inherent in traditional human computational models. The CNN operates by analyzing individual

7 pixel values within the image, undergoing multiple convolution and pooling layers, while

8 incorporating diverse weights, bias elements, and activation functions. This intricate process

9 establishes a mapping between the material's microstructure image and the ultimate prediction

10 outcomes. It is noteworthy that, with an ample abundance of training data, the artificial

11 categorization of microstructures (particles, fibers, various shapes, etc.) holds no bearing on the

12 CNN model's predictive prowess.

33
34

1 4 Conclusion

2 This paper presents a method that uses a convolutional neural network to predict the effective

3 thermal conductivity of fibrous and particulate materials. To generate the digital microstructures of

4 these materials, the QSGS and RGGM methods are employed. The effective thermal conductivities

5 of the x and y axes for each representative volume element are then calculated using Lattice

6 Boltzmann Method. The microstructure models and their effective thermal conductivities are used

f
7 to train the CNN model, which is then optimized by selecting an appropriate number of samples and

oo
8 training parameters. The resulting CNN model is validated with experimental data, demonstrating

r
9
-p
its ability to accurately predict the effective thermal conductivity of materials by capturing their
re
10 microstructure features.
lP

11 The trained CNN model can predict accurately the effective thermal conductivities of
na

12 anisotropic materials along two directions, namely parallel fibers in fibrous materials and aligned
ur

13 elliptical inclusions in particulate materials. Compared to analytical methods, the CNN model
Jo

14 demonstrates outstanding accuracy and efficiency. The CNN model is also used to extrapolate the

15 effective thermal conductivities of hybrid fibrous and particulate models, as well as SEM images

16 from real materials. The predicted results are in close agreement with those obtained by LBM, but

17 with much less computational times. This study actually presents a generic workflow that links

18 material microstructure to its macroscopic physical properties. The training datasets for a CNN

19 model designed to predict the effective thermal conductivity of composite materials are generated

20 using the LBM method. However, it is also feasible to train the CNN model directly with

21 experimental data. Moreover, the CNN can predict any other macroscopic properties, such as

22 Young’s modulus, hardness, etc., that can be correlated to the material's microstructure.
34
35

1 Overall, the proposed method combines microstructural image reconstruction, macroscopic

2 property calculation using the Lattice Boltzmann method, and deep learning techniques to

3 accurately predict the effective thermal conductivity of materials. This work demonstrates the

4 exceptional performance of artificial intelligence algorithms in understanding the microscopic

5 characteristics of materials and linking them to their macroscopic properties. This approach has the

6 potential to accelerate the development of materials design within the framework of the Material

7 Genome Initiative.

f
oo
8 Acknowledgements

r
-p
9 H.F. Zhao acknowledges the support from China Manned Space Engineering Program and the
re
Innovative Research Fund of Chinese Academy of Sciences through the Grant No. JCPYJJ-22002.
lP

10
na

11 Declaration of Competing Interest


ur

12 The authors declare that they have no known competing financial interests or personal
Jo

13 relationships that could have appeared to influence the work reported in this paper.

14 Data availability statement

15 The datasets generated and supporting the finds of this article are available from the

16 corresponding author upon request.

17 References

18 Arenas, J.P., Crocker, M.J., 2010. Recent Trends in Porous Sound-Absorbing Materials. Sound &
19 vibration.
20 Bi, C., Tang, G.H., Hu, Z.J., 2014. Heat conduction modeling in 3-D ordered structures for
21 prediction of aerogel thermal conductivity. Int J Heat Mass Transf 73, 103–109.
35
36

1 https://doi.org/10.1016/j.ijheatmasstransfer.2014.01.058
2 Bostanabad, R., Zhang, Y., Li, X., Kearney, T., Brinson, L.C., Apley, D.W., Liu, W.K., Chen, W.,
3 2018. Computational microstructure characterization and reconstruction: Review of the state-
4 of-the-art techniques. Prog Mater Sci 95, 1–41. https://doi.org/10.1016/j.pmatsci.2018.01.005
5 Chiavazzo, E., Asinari, P., 2010. Reconstruction and modeling of 3D percolation networks of carbon
6 fillers in a polymer matrix. International Journal of Thermal Sciences 49, 2272–2281.
7 https://doi.org/10.1016/j.ijthermalsci.2010.07.019
8 Chollet, F., 2021. Deep learning with Python. Simon and Schuster.
9 Choy, T.C., 2015. Effective Medium Theory: Principles and Applications. Oxford University Press.
10 https://doi.org/10.1093/acprof:oso/9780198705093.001.0001
11 Clyne, T.W., Hull, D., 2019. An Introduction to Composite Materials. https://doi.org/DOI:
12 10.1017/9781139050586
13 Demuth, C., Mendes, M.A.A., Ray, S., Trimis, D., 2014. Performance of thermal lattice Boltzmann

f
14 and finite volume methods for the solution of heat conduction equation in 2D and 3D

oo
15 composite media with inclined and curved interfaces. Int J Heat Mass Transf 77, 979–994.
16 https://doi.org/10.1016/j.ijheatmasstransfer.2014.05.051

r
17 Fu, Z., Corker, J., Papathanasiou, T., Wang, Y., Zhou, Y., Madyan, O.A., Liao, F., Fan, M., 2022.
18
19
-p
Critical review on the thermal conductivity modelling of silica aerogel composites. Journal of
Building Engineering 57. https://doi.org/10.1016/j.jobe.2022.104814
re
20 Gulli, A., Pal, S., 2017. Deep learning with Keras. Packt Publishing Ltd.
21 Hashin, Z., Shtrikman, S., 1963. A variational approach to the theory of the elastic behaviour of
lP

22 multiphase materials. J Mech Phys Solids 11, 127–140.


23 https://doi.org/https://doi.org/10.1016/0022-5096(63)90060-7
na

24 He, J., Li, X., Su, D., Ji, H., Wang, X., 2016. Ultra-low thermal conductivity and high strength of
25 aerogels/fibrous ceramic composites. J Eur Ceram Soc 36, 1487–1493.
https://doi.org/10.1016/j.jeurceramsoc.2015.11.021
ur

26
27 He, Y.-L., Xie, T., 2015. Advances of thermal conductivity models of nanoscale silica aerogel
Jo

28 insulation material. Appl Therm Eng 81, 28–50.


29 https://doi.org/10.1016/j.applthermaleng.2015.02.013
30 Hill, R., 1963. Elastic properties of reinforced solids: Some theoretical principles. J Mech Phys
31 Solids 11, 357–372. https://doi.org/https://doi.org/10.1016/0022-5096(63)90036-X
32 Ishizaki, K., Komarneni, S., Nanko, M., 2013. Porous Materials: Process technology and
33 applications. Springer science & business media.
34 Jabbari, F., Rajabpour, A., Saedodin, S., 2017. Thermal conductivity and viscosity of nanofluids: A
35 review of recent molecular dynamics studies. Chem Eng Sci 174, 67–81.
36 https://doi.org/10.1016/j.ces.2017.08.034
37 Jagjiwanram, Singh, R., 2004. Effective thermal conductivity of highly porous two-phase systems.
38 Appl Therm Eng 24, 2727–2735. https://doi.org/10.1016/j.applthermaleng.2004.03.010
39 Justin, J.-F., Julian-Jankowiak, A., Guérineau, V., Mathivet, V., Debarre, A., 2020. Ultra-high
40 temperature ceramics developments for hypersonic applications. CEAS Aeronaut J 11, 651–
41 664. https://doi.org/10.1007/s13272-020-00445-y
42 Ju, Y., Li, S., Yuan, X., Cui, L., Godfrey, A., Yan, Y., Cheng, Z., Zhong, X., Zhu, J., 2021. A macro-
43 nano-atomic-scale high-throughput approach for material research. Sci Adv 7, eabj8804.
44 https://doi.org/10.1126/sciadv.abj8804

36
37

1 Khan, T., Hameed Sultan, M.T. Bin, Ariffin, A.H., 2018. The challenges of natural fiber in
2 manufacturing, material selection, and technology application: A review. Journal of
3 Reinforced Plastics and Composites 37, 770–779. https://doi.org/10.1177/0731684418756762
4 Kim, K., Lee, Haeun, Kang, M., Lee, G., Jung, K., Kharangate, C.R., Asheghi, M., Goodson, K.E.,
5 Lee, Hyoungsoon, 2022. A machine learning approach for predicting heat transfer
6 characteristics in micro-pin fin heat sinks. Int J Heat Mass Transf 194.
7 https://doi.org/10.1016/j.ijheatmasstransfer.2022.123087
8 Kim, Y.-J., Tan, Y.-F., Kim, S., 2017. Two-dimensional lattice Boltzmann modeling for effective
9 thermal conductivity in carbon black filled composites. J Compos Mater 52, 2047–2053.
10 https://doi.org/10.1177/0021998317737830
11 Kingma, D., Ba, J., 2014. Adam: A Method for Stochastic Optimization. International Conference
12 on Learning Representations. https://doi.org/10.48550/arXiv.1412.6980
13 Larson, N.M., Zok, F.W., 2018. In-situ 3D visualization of composite microstructure during

f
14 polymer-to-ceramic conversion. Acta Mater 144, 579–589.

oo
15 https://doi.org/10.1016/j.actamat.2017.10.054
16 Liu, P.S., Chen, G.F., 2014. Porous Materials: Processing and Applications. Porous Materials:

r
17 Processing and Applications.
18
19
-p
Liu, X., Tian, S., Tao, F., Du, H., Yu, W., 2021. Machine learning-assisted modeling of composite
materials and structures: a review. AIAA Scitech 2021 Forum. https://doi.org/10.2514/6.2021-
re
20 2023
21 Li, X., Liu, Z., Cui, S., Luo, C., Li, C., Zhuang, Z., 2019. Predicting the effective mechanical
lP

22 property of heterogeneous materials by image based modeling and deep learning. Comput
23 Methods Appl Mech Eng 347, 735–753. https://doi.org/10.1016/j.cma.2019.01.005
na

24 Li, Y., Liu, Y., Luo, S., Wang, Z., Wang, K., Huang, Z., Zhao, H., Jiang, L., 2020. Neural network
25 model for correlating microstructural features and hardness properties of nickel-based
superalloys. Journal of Materials Research and Technology 9, 14467–14477.
ur

26
27 https://doi.org/10.1016/j.jmrt.2020.10.042
Jo

28 Ma, J., Nan, C., 2014. Effective-Medium Approach to Thermal Conductivity of Heterogeneous
29 Materials. Annual Review of Heat Transfer 17, 303–331.
30 https://doi.org/10.1615/AnnualRevHeatTransfer.2014007088
31 Michelucci, U., 2019. Advanced applied deep learning: convolutional neural networks and object
32 detection. Springer.
33 Mohamad, A.A., 2019. Lattice Boltzmann Method: Fundamentals and Engineering Applications
34 with Computer Codes. Springer. https://doi.org/https://doi.org/10.1007/978-1-4471-7423-3
35 Mohammadian, M., Jafarzadeh Kashi, T.S., Erfan, M., Soorbaghi, F.P., 2018. Synthesis and
36 characterization of silica aerogel as a promising drug carrier system. J Drug Deliv Sci Technol
37 44, 205–212. https://doi.org/10.1016/j.jddst.2017.12.017
38 Moore, A.L., Shi, L., 2014. Emerging challenges and materials for thermal management of
39 electronics. Materials Today 17, 163–174. https://doi.org/10.1016/j.mattod.2014.04.003
40 Nguyen, D., Tao, L., Ye, H., Li, Y., 2023. Machine learning-based prediction for single-cell
41 mechanics. Mechanics of Materials 180. https://doi.org/10.1016/j.mechmat.2023.104631
42 Nielsen, M.A., 2015. Neural networks and deep learning.
43 Norouzi, S., Kianfar, A., Fakhrabadi, M.M.S., 2020. Multiscale simulation study of anisotropic
44 nanomechanical properties of graphene spirals and their polymer nanocomposites. Mechanics

37
38

1 of Materials 145. https://doi.org/10.1016/j.mechmat.2020.103376


2 O’Shea, K., Nash, R., 2015. An Introduction to Convolutional Neural Networks. ArXiv e-prints.
3 Otsu, N., 1979. A Threshold Selection Method from Gray-Level Histograms. IEEE Trans Syst Man
4 Cybern 9, 62–66. https://doi.org/10.1109/TSMC.1979.4310076
5 Prechelt, L., 2012. Early Stopping — But When?, in: Montavon, G., Orr, G.B., Müller, K.-R. (Eds.),
6 Neural Networks: Tricks of the Trade: Second Edition. Springer Berlin Heidelberg, Berlin,
7 Heidelberg, pp. 53–67. https://doi.org/10.1007/978-3-642-35289-8_5
8 Qian, Z., Wang, Z., Chen, Y., Tong, S., Ge, M., Zhao, N., Xu, J., 2018. Superelastic and ultralight
9 polyimide aerogels as thermal insulators and particulate air filters. J Mater Chem A Mater 6,
10 828–832. https://doi.org/10.1039/c7ta09054d
11 Rabbani, A., Babaei, M., 2019. Hybrid pore-network and lattice-Boltzmann permeability modelling
12 accelerated by machine learning. Adv Water Resour 126, 116–128.
13 https://doi.org/10.1016/j.advwatres.2019.02.012

f
14 Shojaeefard, M.H., Molaeimanesh, G.R., Nazemian, M., Moqaddari, M.R., 2016. A review on

oo
15 microstructure reconstruction of PEM fuel cells porous electrodes for pore scale simulation.
16 Int J Hydrogen Energy 41, 20276–20293. https://doi.org/10.1016/j.ijhydene.2016.08.179

r
17 Srivastava, N., Hinton, G., Krizhevsky, A., Sutskever, I., Salakhutdinov, R., 2014. Dropout: a simple
18
19
-p
way to prevent neural networks from overfitting. J. Mach. Learn. Res. 15, 1929–1958.
Vasudevan, R., Pilania, G., Balachandran, P. V, 2021. Machine learning for materials design and
re
20 discovery. J Appl Phys 129. https://doi.org/10.1063/5.0043300
21 Wang, H., Yin, Y., Hui, X.Y., Bai, J.Q., Qu, Z.G., 2020. Prediction of effective diffusivity of porous
lP

22 media using deep learning method based on sample structure information self-amplification.
23 Energy and AI 2. https://doi.org/10.1016/j.egyai.2020.100035
na

24 Wang, J., Wang, M., Li, Z., 2007. A lattice Boltzmann algorithm for fluid–solid conjugate heat
25 transfer. International Journal of Thermal Sciences 46, 228–234.
https://doi.org/10.1016/j.ijthermalsci.2006.04.012
ur

26
27 Wang, M., Feng, Junzong, Jiang, Y., Zhang, Z., Feng, Jian, 2018. Preparation and properties of the
Jo

28 multi-layer aerogel thermal insulation composites. Heat and Mass Transfer 54, 2793–2798.
29 https://doi.org/10.1007/s00231-018-2320-8
30 Wang, M., Wang, J., Pan, N., Chen, S., 2007. Mesoscopic predictions of the effective thermal
31 conductivity for microscale random porous media. Phys Rev E Stat Nonlin Soft Matter Phys
32 75, 36702. https://doi.org/10.1103/PhysRevE.75.036702
33 Wei, H., Zhao, S., Rong, Q., Bao, H., 2018. Predicting the effective thermal conductivities of
34 composite materials and porous media by machine learning methods. Int J Heat Mass Transf
35 127, 908–916. https://doi.org/10.1016/j.ijheatmasstransfer.2018.08.082
36 Xie, T., He, Y.-L., 2016. Heat transfer characteristics of silica aerogel composite materials: Structure
37 reconstruction and numerical modeling. Int J Heat Mass Transf 95, 621–635.
38 https://doi.org/10.1016/j.ijheatmasstransfer.2015.12.025
39 Xu, A., Shi, L., Zhao, T.S., 2017. Accelerated lattice Boltzmann simulation using GPU and
40 OpenACC with data management. Int J Heat Mass Transf 109, 577–588.
41 https://doi.org/10.1016/j.ijheatmasstransfer.2017.02.032
42 Yang, M., Li, X., Yuan, J., Wen, Z., Kang, G., 2022. A comprehensive study on the effective thermal
43 conductivity of random hybrid polymer composites. Int J Heat Mass Transf 182.
44 https://doi.org/10.1016/j.ijheatmasstransfer.2021.121936

38
39

1 Zeiler, M.D., Fergus, R., 2014. Visualizing and Understanding Convolutional Networks, in: Fleet,
2 D., Pajdla, T., Schiele, B., Tuytelaars, T. (Eds.), Computer Vision – ECCV 2014. Springer
3 International Publishing, Cham, pp. 818–833.
4 Zhai, S., Zhang, P., Xian, Y., Zeng, J., Shi, B., 2018. Effective thermal conductivity of polymer
5 composites: Theoretical models and simulation models. Int J Heat Mass Transf 117, 358–374.
6 https://doi.org/10.1016/j.ijheatmasstransfer.2017.09.067
7 Zhao, H.F., Hu, G.K., Lu, T.J., 2006. Cross-property relations for two-phase planar composites.
8 Comput Mater Sci 35, 408–415. https://doi.org/10.1016/j.commatsci.2005.03.008
9 Zhou, F., Cheng, G., 2014. Lattice Boltzmann model for predicting effective thermal conductivity
10 of composite with randomly distributed particles: Considering effect of interactions between
11 particles and matrix. Comput Mater Sci 92, 157–165.
12 https://doi.org/10.1016/j.commatsci.2014.05.039

13

f
r oo
-p
re
lP
na
ur
Jo

39
-Convolutional neural network (CNN) - a deep learning technique is developed to predict
effective thermal conductivity of fibrous and particulate materials by their microstructural
images.

- By comparing with the results of experiments, analytical models, and computational methods,
CNN model exhibits distinctive "comprehension capabilities" by capturing the scattering
characteristics of heterogeneous materials through artificial intelligence.

-CNN model can predict the novel microstructures that are not included in the training set, as
well as, SEM images from real materials. This highlights the potential of machine learning to
advance materials science and accelerate development of materials with desired properties.

f
r oo
-p
re
lP
na
ur
Jo
Chengcheng Shen:, Methodology, Software, Validation, Investigation, Writing- Original draft
preparation, Visualization, Data Curation, Formal analysis.
Qiang Sheng: Resources, Investigation, Project administration.
Haifeng Zhao: Conceptualization, Investigation, Formal analysis, Resources, Writing - Review
& Editing, Visualization, Supervision, Funding acquisition.

f
r oo
-p
re
lP
na
ur
Jo
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐ The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

of
ro
-p
re
lP
na
ur
Jo

You might also like