Download as pdf or txt
Download as pdf or txt
You are on page 1of 89

Journal of Electronic Materials

A computational multiscale modeling method for nanosilver sintered joints with


stochastically distributed voids
--Manuscript Draft--

Manuscript Number: JEMS-D-23-01855R1

Full Title: A computational multiscale modeling method for nanosilver sintered joints with
stochastically distributed voids

Article Type: Original Research

Keywords: Nanosilver sintered joints; Stochastically distributed voids; Multiscale modeling;


Monte Carlo

Abstract: High power density is required in wide band gap power semiconductor packaging,
which has led to the popularity of sintered nanosilver as an interconnecting material.
However, affected by stochastically distributed voids in its microstructure, this material
practically exhibits instability leading to reduced reliability. In this paper, a
computational multiscale modeling method is originally proposed to simulate the
influence of micro voids on macro properties, providing an efficient tool to analyze the
aforementioned problem. At the micro-scale, the three-parameter Weibull distribution of
the equivalent Young’s modulus and the Normal distribution of the equivalent
Poisson’s ratio are captured by Monte Carlo-based finite element simulation on the
reconstructed stochastic representative elements, where the density and distribution
morphology of micro voids are taken into consideration. At the macro-scale, the effect
of the microscopic voids is transferred through a random sampling process to construct
the multiscale model. The effectiveness and validity of the proposed method are
verified through experimental case studies involving the modeling of nanosilver
sintered joints sintered at temperatures of 275°C and 300°C. Besides, the effects of the
sintering temperature on the dispersion of the micro voids, the distribution fluctuation of
the constitutive parameters, and the mechanical properties are also discussed based
on numerical and experimental results.

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Cover letter Click here to access/download;Cover letter;Cover Letter-
revised.pdf

AAU Energy, Aalborg University


Aalborg, Denmark
19/Dec./2023
Professor Jamie Buonato
Editor
Journal of Electronic Materials

Dear Prof. Jamie Buonato,

I am writing to submit the revised version of our manuscript titled A computational multiscale modeling method
for nanosilver sintered joints with stochastically distributed voids " for consideration in the Journal of Electronic
Materials. We deeply appreciate the valuable feedback provided by the reviewers and the editorial office and have
made comprehensive revisions to address all the comments and suggestions.

Key revisions and responses include:

1. Enhanced Modeling and Simulation Details: We elaborated on our original computational multiscale
modeling method, emphasizing its novelty and detailing the steps taken to ensure accurate simulation
results. Sections detailing the methodology and algorithms used have been thoroughly revised for clarity.
2. Additional Data and Comparisons: Recognizing the need for more comprehensive data, we have
included new comparisons with literature-reported shear tests and expanded our discussion on the
influence of sintering temperature. Although unable to conduct additional experiments due to institutional
changes, we believe these additions significantly bolster our findings.
3. Image and Figure Enhancements: Following the guidance for higher resolution and clarity, we have
uploaded high-resolution images (300 or 600 dpi) in TIF format. Scale bars have been added to all
micrographs as required, and the graphical abstract has been included at the beginning of the manuscript.
4. Formatting Adjustments: We have reformatted the manuscript thoroughly and made sure that it meets
the specified standards for readability and quality.

Each of these points has been thoroughly addressed in our “Point-to-Point” response included in the Appendix
of the Response to Reviewers. The revised manuscript, along with the supplementary LaTeX files, has been uploaded
to the Editorial Manager system. We have ensured that all changes are highlighted in the manuscript for an efficient
review process.

We believe that these revisions have significantly improved our manuscript and address all the concerns raised
by the reviewers and the editorial office. We are confident that our study will make a valuable contribution to the
field and will be of interest to the readers of the Journal of Electronic Materials.

Thank you for considering our revised manuscript for publication in the Journal of Electronic Materials. We
eagerly await your response and are happy to provide any further information or clarification you may need.

Yours sincerely,

Wendi Guo

Corresponding author,

E-mail: wg@energy.aau.dk
Response to Reviewer Comments

AAU Energy
Aalborg, Denmark
19/Dec./2023
Prof. Jamie Buonato
Editor
Journal of Electronic Materials

Dear Prof. Jamie Buonato,

Subject: Submission of Revised Manuscript JEMS-D-23-01855


Manuscript Title: A computational multiscale modeling method for nanosilver sintered joints with
stochastically distributed voids

First of all, the authors would like to express our deepest gratitude to you and the anonymous
reviewers for your invaluable and constructive feedback on our manuscript. Your insights have been
instrumental in enhancing the quality and depth of our paper. We are now submitting a revised version,
which has been updated considering the comments received.

In the Appendix, we provide detailed "Point-to-Point" responses to each specific comment made by
the reviewers. To facilitate an efficient review, we have highlighted the revisions in the manuscript with
a yellow background. We hope that these revisions address the concerns raised and further strengthen
our work.

We would be grateful if you could review the revised manuscript and provide any additional
feedback you deem necessary.

Thank you once again for your valuable time and input.

Wendi Guo
Corresponding author
E-mail: wg@energy.aau.dk
Appendix: “Point-to-Point” responses to the specific comments

Reviewer 1

1. Exploring the nano silver sintered joints in high-power devices is a compelling subject.
Nevertheless, the manuscript currently focuses solely on Finite Element (FE) modeling with
experimental parameters, such as porosity. Unfortunately, this approach hinders the model's ability to
accurately predict the behaviors of sintered joints. It is recommended that the author conducts
additional simulations to predict the shear force of the interfacial joints and compares these results
with shear test outcomes.

Response:

Thank you for your insightful comments on our manuscript. We acknowledge the importance of
thorough and precise modeling in understanding the behavior of nanosilver sintered joints in high-
power devices.

In addressing your concerns about the scope of our Finite Element (FE) modeling, we wish to
emphasize that our study extends beyond analyzing the effects of porosity on sintered joints. In the
revised Section 2.2, titled "Reconstruction of Voids," we detail our comprehensive consideration of
void distribution characteristics, including size, dispersion, and shape. To emphasize the void
characterization parameters involved in our reconstruction process, we have updated and highlighted
Figure 2 and the reconstruction process in the manuscript.

Our multiscale modeling and simulation strategy commenced at the microscale, where we analyzed
Scan Electron Microscopy (SEM) images of actual sintered joints. Employing threshold segmentation
and contour recognition techniques, we successfully extracted the aforementioned void distribution
characteristics. Based on this, we reconstructed stochastic representative elements with similar void
dispersion characteristics using the random medium theory and similarity analysis, facilitating
subsequent simulation analyses.

Furthermore, at the microscale, we employed Monte Carlo-based FE simulations to analyze the impact
of void presence and their dispersion on Young's modulus and Poisson's ratio of sintered joints. We
also utilized parameter fitting and resampling methods to transfer these effects from the microscale to
the macroscale behavior of nanosilver sintered joints. On the macroscale, we computed the shear
strength of the nanosilver sintered samples using FE simulation analysis, aligning with your
recommendation to include experimental comparisons. These comparisons affirm the efficacy of our
multiscale computational model in predicting the mechanical properties of nanosilver sintered joints.

To illustrate the ability of the proposed model to accurately predict the behaviors of sintered joints,
we have also included a comparison with shear tests reported in the literature. We examined the
influence of sintering temperature on the sintering process, analyzing both microscale and macroscale
results from our simulations. These additional insights are detailed in Section 4.4, which has been
thoroughly revised and highlighted for ease of review.

We believe these updates and clarifications will address your concerns and enhance the manuscript's
contribution to the field.
2. The absence of solder paste utilization in the study needs correction by the author.

Response:

Thank you for pointing out the omission regarding the use of solder paste in our study. We appreciate
your keen observation and the opportunity to rectify this aspect of our research.

In response to your comment, we have enriched Section 3.1, titled "Sample Manufacture," with
detailed information about the nanosilver sintered solder paste used in our experiments. This
enhancement includes a comprehensive description of the solder paste's composition, viscosity, and
moisture absorption temperature, along with the specific methods employed in its preparation.

Additionally, we have updated Figure 4, now titled "The Manufacturing Process of Nanosilver
Sintered Joint Samples." This revised figure clearly illustrates how the nanosilver sintered solder paste
was utilized in the manufacturing of our samples. We have made sure to highlight this aspect
prominently to provide a clear visual representation of its application in our experimental process.

We trust that these amendments and additions comprehensively address the concern you raised and
further enhance the clarity and completeness of our manuscript.

3. The SEM images require resolution enhancement and the inclusion of a scale for improved clarity
and accuracy.

Response:
Thank you for your valuable feedback concerning the SEM images in our manuscript.

Following your suggestion, we have updated Figures 6 and 8 to clearly indicate the scale bars.
Additionally, to further support the quality of our visual representations, we have submitted the high-
resolution source files of these images.
Reviewer 2

1. This manuscript presents a multiscale modeling method for predicting the mechanical property of
nanosilver sintered joints. The authors ought to emphasize, first of all, whether this method is invented
by themselves or bought from somewhere else, or to be specific on what are their contributions for
this method. Meanwhile, the finite element code and other algorithms that they have used to produce
the results for this manuscript should be clearly written in details. The current version of the
manuscript tried to play vague on those issues using quite general terms e.g. "using OpenCV-based
image processing method", "FE modeling (here taking ANSYS as an instance)".

Response:
Thank you for your insightful feedback and for raising important questions regarding the methodology
of our study on the multiscale modeling of nanosilver sintered joints.

Regarding the Originality of Our Method:

We would like to clarify that the computational multiscale modeling method presented in this
manuscript is indeed an original contribution developed by our research team. Our method is designed
to analyze the impact of stochastically distributed micro voids on the macroscopic properties of
nanosilver sintered joints. This approach is both operable and robust, representing a significant
advancement in this field.

At the microscale, we have built upon a characterization system for micro-voids in nanosilver sintered
joints, which was initially introduced in our previously published paper [1]. This involves the use of
random medium theory and threshold segmentation methods for reconstruction. In this manuscript,
we provide a concise introduction to these foundational methods.

Innovatively, our current research introduces a method for model conversion using text files and Finite
Element (FE) modeling using ANSYS Parametric Design Language (APDL), ensuring perfect
alignment between the FE model and the stochastic representative element (SRE). We have also
developed a Monte Carlo FE simulation approach to analyze how the dispersion of micro-voids affects
the distribution of mechanical properties such as Young's modulus and Poisson's ratio of the sintered
joints. Furthermore, we propose a scale transition scheme based on parameter distribution resampling
to convey the impact of micro void distribution characteristics. Through these methodologies, we have
successfully established a computational multiscale modeling approach.

In the revised manuscript, we have made edits to sections including the Abstract, Introduction, and
Conclusion to clearly articulate the originality of our method.

Regarding Specifics of Finite Element Code and Algorithms:

Addressing the second aspect of your question, we have detailed the algorithms used in our study in
Section 2, "Methodology." Additionally, to provide comprehensive insight into our process, we have
included the core codes used in our simulations in the Appendix of the manuscript. This ensures
transparency and allows for a thorough understanding of the technical underpinnings of our research.
Besides, the codes are provided as a supplementary document for easy access.

[1] Qian, C., Sun, Z., Fan, J., Ren, Y., Sun, B., Feng, Q., Yang, D., Wang, Z.: Characterization and reconstruction for
stochastically distributed void morphology in nano-silver sintered joints. Materials and Design 196 (2020)
https://doi.org/10.1016/j.matdes.2020.109079
2. Needs at least one more data point: the data (experimental and modeling) for one more temperature
inbetween is required to verify the results presented in the manuscript.

Response:
Thank you for your suggestion to include additional data points for verifying the results presented in
our manuscript. We understand the importance of comprehensive data to strengthen the validity of our
findings.

Regrettably, due to changes in our research institution, we are currently unable to prepare more
samples or conduct additional experiments and simulations at varying temperature conditions to
further demonstrate the robustness of our method. We acknowledge this limitation and apologize for
the inconvenience.

To address this concern, we have made significant revisions to Section 4.4 of our manuscript. In this
section, we have included a new comparison with shear test results from the literature [2] and [3]. The
experimental samples selected from the literature were prepared under the same sintering temperatures
and pressures consistent with our sintering process. Notably, the literature [3] reports a shorter
sintering duration of only 5 minutes at a temperature of 300°C, resulting in slightly lower results
compared to our experimental and simulation outcomes.

Overall, the FE simulation analysis conducted based on our proposed multiscale method demonstrates
consistency with both our experimental results and those reported in the literature, thus affirming the
feasibility and accuracy of our approach. We have also analyzed the sources of simulation error,
indicating potential areas for further refinement in future applications of our model.

Additionally, we have conducted a qualitative analysis of the sintering process mechanism to


understand the impact of sintering temperature on the performance of sintered joints. Our theoretical
conclusions align with our simulation results at both the microscale and macroscale levels. We found
that an increase in sintering temperature enhances the joint's resistance to plastic deformation at the
macroscale by promoting grain boundary formation. Moreover, at the microscale, a higher sintering
temperature contributes to an increase in the Young's modulus of the sintered joints.

We believe these additions and analyses in our revised manuscript effectively address the data point
concern and further substantiate the validity of our research findings.
[2] Lei, T.G., Calata, J.N., Lu, G.Q., Chen, X., Luo, S.: Low-temperature sintering of nanoscale silver paste for attaching
large-area (>100 mm2) chips. IEEE Transactions on Components and Packaging Technologies 33(1), 98–104 (2010) 28
https://doi.org/10.1109/TCAPT.2009.2021256
[3] Ide, E., Angata, S., Hirose, A., Kobayashi, K.F.: Metal-metal bonding process using Ag metallo-organic nanoparticles.
Acta Materialia 53(8), 2385–2393 (2005) https://doi.org/10.1016/j.actamat.2005.01.047

3. All the microstructural images need scale bars.

Response:

Thank you for your valuable feedback regarding the inclusion of scale bars in our microstructural
images. In response to your comment, we have updated Figures 6 and 8, which depict the
microstructure of the sintered joints. These figures now include marked scale bars, providing a
reference for size and aiding in the accurate analysis of the microstructures presented.
4. Fig. 11 is a bit confusing: why the displacement results from SRE are not quite related to the voids
morphology and distribution? What are the properties of voids used for the displacement calculation?

Response:
Thank you for your observations regarding Figure 11 in our manuscript. We appreciate your keen
attention to detail and the opportunity to clarify these aspects of our study.

Addressing the Displacement Results from SRE:

We acknowledge that the displacement results depicted in Figure 11 for two different sintering
temperatures might have caused some confusion due to the variation in the numerical range of the
color bars. To rectify this and enhance clarity, we have explicitly specified the maximum displacement
values for each simulation result in Section 4.1. This adjustment should facilitate a more
straightforward comparison and understanding of the displacement results under different conditions.
Additionally, the color bars in Figure 11 are enhanced to provide a clearer vision.

Influence of Voids on Displacement:

Regarding the impact of voids on displacement, if the nanosilver sintered SREs are homogenous and
void-free, the displacement isoclines would be parallel and perpendicular to the loading direction.
However, the actual results reveal that the presence of voids leads to non-uniform displacement
variations. Voids with a larger aspect ratio along the loading direction tend to cause a more significant
displacement gradient. Additionally, the sparser the distribution of voids, the more pronounced the
displacement response.

Properties of Voids in Displacement Calculation:

In our microscale simulations to calculate displacement under tensile stress and thereby derive
equivalent Young's modulus and Poisson's ratio, the voids are considered as gaseous phases.
Therefore, in the solid mechanics calculations, the physical parameters of the voids are not included.
In the modeling of the SRE, we excluded the void portion and modeled only the sintered nanosilver
material. We have made modifications to Subsection 2.3.1 to emphasize this particular aspect of our
modeling approach.

We believe these clarifications and revisions in our manuscript will resolve the confusion surrounding
Figure 11 and provide a more comprehensive understanding of the displacement results and their
relationship to void morphology and distribution.
Editorial office:
1. Please upload high-resolution (at least 300 dpi) jpg, tif, or eps files for each figure separately in
Editorial Manager. Do not upload your figures in a Word document file.

Response:
Thank you for your instructions regarding the figure uploads. We have complied with the requirements
and uploaded each figure as a separate TIF file, with resolutions of either 300 or 600 dpi, ensuring
high quality and clarity.
2. Please upload the manuscript as a Word file. If you have only a LaTeX file and cannot upload a
Word file, please upload a PDF as the “Manuscript” file but upload the LaTeX file(s) as a
Supplementary file(s).

Response:
Thank you for your guidance on manuscript submission formats. We have prepared our manuscript
using LaTeX. According to your comments and further notification, we have uploaded the .tex files
as the Manuscript and the PDF version as the Supplementary File for reference.
3. Please include your graphical abstract at the beginning of your manuscript, after the title page.
Additionally, we ask that you upload the graphical abstract into the system as a separate, high-
resolution (at least 300 dpi) figure file (jpg, tif, eps, or png). Please label the graphical abstract clearly
in the system.

Response:
Thank you for the instructions regarding the inclusion of a graphical abstract. We have incorporated
the graphical abstract in our manuscript. Additionally, we have uploaded a separate high-resolution
file of the graphical abstract in TIF format to the system, labeled as the graphical abstract.
4. Please use SI units for hours (h), minutes (min), and seconds (s).

Response:
We have made the necessary changes to units in our manuscript, and these changes have been
highlighted in the manuscript for easy identification and review.
5. Change tables from Arabic numbers to Roman numerals. For example, change Table 1 to Table I.

Response:
We have made the necessary changes to the numbering of the tables in our manuscript, and these
changes have been highlighted in the manuscript for easy identification and review.
6. Figures 6, 8: All micrographs must have scale markers. Scale markers and units on micrographs
should be placed in a white box within the image using a thicker line for the scale and a larger font
size used for the units for better readability.

Response:
In response to your request, we have updated Figures 6 and 8 in our manuscript to include scale
markers. These modifications have been highlighted in the manuscript for ease of identification and
review.
Manuscript Click here to access/download;Manuscript;Manuscript.tex

Click here to view linked References

1
2
3
4
5
6 A computational multiscale modeling method
7
8 for nanosilver sintered joints with stochastically
9
10 distributed voids
11
12
13 Zhongchao Sun1 , Wendi Guo1*, Asger Bjørn Jørgensen1
14 1* AAU
15 Energy, Aalborg University, Pontoppidanstræde 101, Aalborg,
16 9220,Denmark.
17
18
19 *Corresponding author(s). E-mail(s): wg@energy.aau.dk;
20 Contributing authors: zs@energy.aau.dk; abj@energy.aau.dk;
21
22
23
24
High power density is required in wide band gap power semiconductor packaging,
25
which has led to the popularity of sintered nanosilver as an interconnecting mate-
26
27 rial. However, affected by stochastically distributed voids in its microstructure, this
28 material practically exhibits instability leading to reduced reliability. In this paper,
29 a computational multiscale modeling method is originally proposed to simulate the
30 influence of micro voids on macro properties, providing an efficient tool to analyze the
31 aforementioned problem. At the micro-scale, the three-parameter Weibull distribu-
32 tion of the equivalent Young’s modulus and the Normal distribution of the equivalent
33 Poisson’s ratio are captured by Monte Carlo-based finite element simulation on the
34 reconstructed stochastic representative elements, where the density and distribution
35 morphology of micro voids are taken into consideration. At the macro-scale, the effect
36 of the microscopic voids is transferred through a random sampling process to construct
37 the multiscale model. The effectiveness and validity of the proposed method are ver-
38 ified through experimental case studies involving the modeling of nanosilver sintered
39 joints sintered at temperatures of 275°C and 300°C. Besides, the effects of the sinter-
40 ing temperature on the dispersion of the micro voids, the distribution fluctuation of
41
the constitutive parameters, and the mechanical properties are also discussed based
42
on numerical and experimental results.
43
44
45
46
47
48
49
50 1
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13 Graphical Abstract
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50 2
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1 Introduction
1
2 Benefiting from high breakdown voltages, high operating temperatures, and high
3 switching frequencies, wide band gap (WBG) semiconductors such as silicon carbide
4 (SiC) and gallium nitride (GaN) are dominating the industry of electronics packaging
5 to ensure high power density [1–3]. To release the potential of WBG semiconductors
6 adequately and reliably, traditional Pb and lead-free solders with lower melting tem-
7 peratures [4, 5] are being substituted by new interconnecting materials. Considering
8 the prerequisites for bonding power chips concerning mechanical connection, electri-
9 cal transportation, and heat dissipation, low temperature [6, 7] sintering nanosilver
10 (Ag-NP) emerges as a compelling option within various new attachment technologies
11 [8, 9], featuring a high operating temperature of up to 961.8°C [10] and low electrical
12
and thermal resistance of 429 W·m−1 K−1 and 1.6 µΩ·cm [11], respectively.
13
During the sintering process, silver nanoparticles are driven by the surface energy
14
15 and defective energy of the sintering paste to coalesce into a bulk solder, where
16 mechanical pressure and an external heat source are commonly involved to obtain a
17 denser structure. However, the applied pressure and excessive temperature may cre-
18 ate defects and cracks on the chip and substrate [12, 13], so they are limited to some
19 extent. Due to the inherent nature of the sintering mechanism, stochastically dis-
20 tributed voids inevitably exist in the microstructure of nanoparticle-sintered joints,
21 subsequently affecting the macro-level mechanical [14], thermal [15], and electrical
22 properties [16]. By experimental methods [17], the micro voids observed by scanning
23 electron microscope (SEM) with larger porosity were found to lead to a decrease in
24 the tensile strength and bonding strength in nanosilver sintered joints of larger joint
25 cross-sections [18] and inadequate thickness [19]. In the thermal and electrical fields, a
26 35% increase in porosity will result in approximate 80% and 25% decreases in thermal
27 conductivity and electrical conductivity of nanosilver sintered joints, respectively [16].
28 Triggering the stress singularity point [20], the micro voids are always related to the
29
crack initiation and propagation, then leading to reliability issues [21, 22]. To model
30
the relationship between the voids and the macro properties, the material functions-
31
32 based mathematical methods [23] and the model-based numerical simulation methods
33 [24–26] are performed. A multiphysics coupling model [27] of a single-chip 3.3kV/50A
34 nanosilver sintered press-pack IGBT simulated the impacts of the void distribution,
35 number, and ratio on the electrical-thermal stress, junction temperature, on-state
36 voltage, and the chip current distribution, while the voids are regarded as perforated
37 circles with a size of 0.05 mm rather than the actual morphology. The numerical model
38 proposed by Fei et al. [28] showed the equivalent thermal conductivity of porously sin-
39 tered silver is affected by the void shape of the circle, ellipse, and rectangle. Another
40 research [29] used Focused-Ion-Beam tomography to get the real sintering structures
41 in the silver film and analyzed their influence on the macroscopic stress state and yield
42 locus of sintered silver by formulating a thermodynamic consistent continuum model,
43 but the dispersion of the stochastically distributed voids is not included.
44 To take a more accurate analysis of nanosilver sintered joints considering the effects
45 of the porous microstructure, the real morphology has to be taken into account rather
46 than relying on simplifications [30, 31]. At the same time, the dispersion of the voids
47
needs to be considered to illustrate the property fluctuation under the same sintering
48
49
50 3
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
process [32]. Facing the problem, the morphology of the stochastically distributed
1 voids is randomly reconstructed by keeping the same dispersion characteristics in our
2 previous research [33], which can be used as the modeling input for the finite element
3 simulation but cannot be used directly. Although in power electronic packaging, the
4 bonding layers usually have small geometrical dimensions (at the millimeter level) [34],
5 the nanosilver sintered joint has a 5000-times geometric difference compared to the
6 microstructure with stochastically distributed voids (at the micron meter level) [35].
7
If modeling the power module containing microscopic voids directly for simulation
8
calculations, it will bring an enormous number of meshes and computational resources,
9
10 despite the improved simulation accuracy.
11 As one popular indirect modeling method, the multi-scale theory [36] is used
12 to analyze the properties of materials such as concrete [37], graphene-reinforced
13 polypropylene (PP /graphene) nanocomposites [38], and duplex stainless steel materi-
14 als [39] at the macroscopic scales by using the representative volume elements (RVEs)
15 to maintain the disturbance from the micro-scale and without modeling all geometric
16 details [40], which all have good agreements with the available test data. Given that
17 nanosilver sintered joints exhibit a similar void morphology and multi-scale feature as
18 the above materials, it is expected that such a theory can also effectively model their
19 properties. However, the RVE only works on periodic elements [41] while the real micro
20 morphologies in nanosilver sintered joints are random, and the properties of the micro
21 porous structure will behave as a stochastic distribution, accordingly. Therefore, the
22 universal multiscale modeling pipeline does not work on nanosilver sintered joints. To
23 solve the problem and build the multiscale model of nanosilver sintered joints which
24 is crucial to analyze the relationship between the micro voids and macro properties,
25
a computational method including Monte Carlo simulation and sampling is originally
26
proposed and verified to construct the multiscale model for nanosilver sintered joints
27
28 with stochastically distributed voids in this paper.
29 The remaining parts of this paper are organized as follows: Section 2 presents
30 detailed information on the proposed approach to model nanosilver sintered joints
31 from the microscale to the macroscale by the computational method. Section 3 intro-
32 duces the manufacturing process of samples and the experimental results of the SEM
33 observation and shear tests. Section 4 provides the application and validation scenarios
34 for the proposed approach and analyzes the influence of the stochastically distributed
35 voids, respectively. In the end, the major conclusions from this study are drawn in
36 Section 5.
37
38
39 2 Methodology
40
41 When a nanosilver sintered joint with stochastically distributed microscopic voids
42 is loaded, voids are not subjected to the load directly, but voids’ density, size and
43 distribution morphology affect the generation of internal local pressure and thus the
44 macroscopic response of the sintered joint. In the case of mechanical loads, this effect is
45
reflected as a change in the macroscopic constitutive relationship of the sintered joint
46
which will show a certain statistical distribution due to the statistical characteristics
47
48 of the arbitrarily distributed microscopic voids.
49
50 4
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
2.1 Overall methodology flow
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19 Fig. 1: The flow diagram of the proposed multiscale modeling method
20
21
22
23
24 The whole procedure of building a finite element analysis (FEA) model of nanosil-
25 ver sintered joints considering the effect of randomly distributed microscopic voids
26 based on a computational multiscale approach is shown in Fig. 1. This proposed
27
procedure includes the following steps:
28
(1) Based on the method of characterizing and reconstructing the random void mor-
29
30 phology of the nanosilver sintered joints, which was previously invented and published
31 in [33] and presented in Section 2.2, reconstruct the stochastic representative elements
32 (SREs) that are able to characterize the microscopic void distribution features of the
33 nanosilver sintered joints.
34 (2) Taking in the SREs reconstructed above, establish corresponding FEA models
35 according to the method in Section 2.3 and execute the Monte Carlo finite element
36 simulations on the micro-scale to obtain the stochastic constitutive model responses
37 from different microscopic void morphologies of the same sintered joint.
38 (3) Count the probability distribution of the constitutive parameters calculated by
39 simulation in the previous step, find the most matched statistical model, and analyze
40 the distribution pattern using the method developed in Section 2.4.
41 (4) Model the macroscopic nanosilver sintered joint in the sequence of 3D mod-
42 eling, meshing, material definition, and boundary conditions application, where the
43 constitutive parameters of each element are sampled from the fitted models in step
44 (3). The detailed procedure is referred to the instructions in Section 2.5. Then the
45
model could be solved in FEA with an ability to consider the influence of microscopic
46
voids on the macroscopic properties of the nanosilver sintered joint without increasing
47
48 the geometric complexity.
49
50 5
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
2.2 Reconstruction of voids
1
2 The flow diagram of reconstructing SREs of the microstructure of nanosilver sintered
3 joints is shown in Fig. 2.
4 (1) Obtain the SEM images recording the microstructure morphology of the
5 nanosilver sintered joints clearly, separate and distinguish the stochastically dis-
6 tributed voids from the sintered nanosilver using OTSU algorithm-based threshold
7 segmentation image processing methods. Extract the porosity and the characteriza-
8 tion parameters matrix, which includes the void area (A), distance (l ’) and deflection
9 angle (α) from the geometric center, and the void aspect ratio (δ), noted as P.
10 (2) Based on the random medium (RM) theory and the porosity-based threshold
11 segmentation algorithm, generate a considerable number of void morphology samples
12 with discrete autocorrelation function parameter inputs, and extract their characteri-
13 zation parameters matrices Qi as step (1) to form a void morphology database, denoted
14 as Q.
15 (3) Compare the similarity between matrix P and Qi by using the Jensen-
16
Shannon (JS) divergence algorithm, find the characterization parameters matrix Qmax
17
with the highest matching degree, and note its input autocorrelation parameters of
18
19 corresponding void morphology as pending reconstruction parameters.
20 (4) In the similarity calculation, if the similarity is less than the specified tolerance,
21 the input parameters are refined with reference to pending reconstruction parameters,
22 and steps (2-3) are repeated until the specified tolerance is satisfied, and the final
23 reconstructed parameters are obtained.
24 (5) Following the same method in step (2), establish void morphology samples
25 with the same void distribution characteristics as the input nanosilver sintered joint
26 microstructure by using the final reconstructed parameters. Since the effect of the
27 stochastic parameter used in RM theory, the samples with the same modeling param-
28 eters show different morphology features, which exhibit statistical randomness. Hence,
29 the samples here are noted as the SREs of the nanosilver sinter joints.
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47 Fig. 2: Stochastic representative elements reconstruction flow diagram
48
49
50 6
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
2.3 Monte Carlo-based Finite element simulation
1
2 2.3.1 FE modeling of SRE
3 The SRE created in Section 2.2 stores the microstructure morphology information of
4 the nanosilver sintered joint in the form of a binarized matrix, and each matrix cell
5 represents a pixel in the bitmap, where 0 represents the pixel for the void and 1 repre-
6
sents the medium. In the geometric modeling stage in the FE simulation environment,
7
a medium pixel in the SRE can be equated to a two-dimensional plane element.
8
9 Based on this principle, the proposed method uses an intermediate file to record
10 the pixelated structure characteristics of the SRE and nondestructively converts the
11 SRE from a numerical model to a geometric model. The specific process is illustrated
12 in Fig. 3 and introduced as follows. The core code run in the Ansys Parametric Design
13 Language (APDL) environment for SRE FE modeling is provided in the Appendix A.
14 (1) Extract the morphology features of SRE in the format of ID, row, column,
15 and feature value, marked as N, I, J, and F, respectively, and export it as a .txt file,
16 which can be recognized by ANSYS Mechanical easily. In this intermediate file for
17 SRE modeling, the ID is used to record the SRE size, the row and column are used to
18 locate the pixel points, and the feature value is used to distinguish whether the pixel
19 characterizes a void or a medium.
20 (2) Import the .txt file into ANSYS Mechanical and calculate the actual size of
21 the SRE in both X and Y directions, which is the product of the maximum values of
22 the row and column and the plotting scale (the actual size of 1 pixel) of the bitmap.
23 Following the size limit, draw the key points used to describe all the plane elements
24
in ANSYS geometry space, where the interval of the key points is the length of the
25
plotting scale.
26
27 (3) Read the pixel feature values in the .txt file, identify the medium compo-
28 nents, and connect the corresponding key points according to their locations to form
29 2D planes. Connect all two-dimensional planes representing the sintered nanosilver
30 material by the Boolean operation to build the SRE geometry model.
31 (4) In the meshing stage, the plotting scale is set as the quadrangular element
32 size to guarantee its success regardless of the morphological complexity induced by
33 stochastically distributed voids. Then, the FE model of SRE is prepared for further
34 simulation.
35
36
37
38
39
40
41
42
43
44
45
46 Fig. 3: The flow to convert the stochastic representative element into finite element
47 model
48
49
50 7
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
2.3.2 Monte Carlo-based FE simulation
1
2 During the elastic deformation phase, the effect of microscopic stochastically dis-
3 tributed voids of the nanosilver sintered joint will be reflected as the variation of the
4 equivalent Young’s modulus and Poisson’s ratio of the SRE by using Hooke’s law to
5 describe the constitutive relationship of the nanosilver sintered joint.
6 With the FE simulation, the tensile deformation in the tensile direction and the
7 transverse deformation in the vertical direction of the SRE subjected to the uniform
8 tensile load can be extracted, and according to Eq. 1 and 2, the equivalent Young’s
9 modulus and Poisson’s ratio can be calculated respectively.
10
11
12 P lx
E= (1)
13 ∆lx
14
15 where the X-direction is regarded as the tensile direction, E in MPa is the equiva-
16 lent Young’s modulus of SRE, P in MPa represents the uniform tensile load, lx in µm
17 is the size of SRE along the tensile direction and ∆l x in µm is the tensile deformation.
18
19
20 εy ∆ly · lx
21 ν=− =− (2)
εx ∆lx · ly
22
23 where ν is the equivalent Poisson’s ratio of SRE, εx and εy are the tensile strain
24
and transverse strain, l y in µm is the size of SRE along the transverse direction and
25
∆l y in µm is the transverse deformation.
26
27 In order to analyze the effect of the randomness of the microscopic void distribution
28 on the constitutive relationship of nanosilver sintered joint, the Monte Carlo-based FE
29 simulation was conducted. By coding an APDL program displayed in the Appendix
30 B, the different SRE of the same nanosilver sintered joint reconstructed in Section 2.2
31 are repeatedly imported to update the FE model and simulated to obtain correspond-
32 ing the equivalent Young’s modulus and equivalent Poisson’s ratio, according to the
33 method proposed in this Section.
34
35 2.4 Parameter fitting
36
37 From the Monte Carlo-based FE simulation responses, different SREs of the same
38 nanosilver sintered joint will result in different equivalent Young’s modulus and Pois-
39 son’s ratio observations, which come from the same stochastic distribution, but are
40 independent of each other.
41 To characterize the microscopic constitutive parameters distribution of the nanosil-
42 ver sintered joint considering the randomness of the microscopic void distribution, the
43 proposed method first calculates the frequency distribution of the discrete constitu-
44 tion parameters observations obtained from the Monte Carlo-based FE simulation in
45
Section 2.3 and plots it in the form of a continuous probability density function, and
46
then takes the commonly used statistical distributions including the normal, lognor-
47
48 mal, exponential, Weibull and three-parameter Weibull distributions to fit that. The
49
50 8
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
distribution form with the highest acceptance probability is selected as the stochas-
1 tic distribution of the constitutive parameters, and the distribution parameters are
2 determined at the same time.
3
4
5 2.5 Sampling and macro modeling
6
7 Considering the size difference between the micro voids and the macro sintered joints
8 and to reduce the computational burden, rather than modeling all the voids in the
9 joints, the proposed macroscopic simulation modeling method based on Monte Carlo-
10 based FE simulation and parameters sampling is developed as below:
11 In the nanosilver sintered joint FE simulation, a macroscopic geometric model
12 matching the actual size of the simulation object is established first, and the element
13 size of the joint is controlled to be the same as that of the SRE in the meshing stage.
14 Considering the propagation of the influence of the microscopic voids on the consti-
15 tutive relationship of the nanosilver sintered joint along the scale growth, the physical
16 parameters of each sintered joint element should be consistent with the microstruc-
17 ture in the macroscopic FE simulation modeling, and considering the randomness
18
of the microscopic void distribution, the constitutive parameters of each element of
19
the sintered joint will be arranged as the random sampling samples of the stochastic
20
21 distribution of the equivalent Young’s modulus and the equivalent Poisson’s ratio.
22 Based on the obtained stochastic distribution forms in Section 2.4, the samples
23 with the same number of elements as the nanosilver sintered joint are generated by a
24 random number-based direct sampling method, and the samples are assigned to each
25 element of the nanosilver sintered joint using the APDL program. With this, the mul-
26 tiscale modeling for nanosilver sintered joints considering the influence of microscopic
27 stochastically distributed voids on macroscopic properties is completed.
28
29
30 3 Experiments
31
32 3.1 Sample manufacture
33
34 The manufacturing process of the nanosilver sintered joint samples with typical "sand-
35 wich" structure is shown in Fig. 4, and the sample components are shown in Table
36 I. The nanosilver paste [42] consists of spherical and quasi-spherical Ag nanoparti-
37 cles with an average particle size of 80 nm and 800 nm and organic matters, where
38 polyethylene glycol is used to adjust the viscosity and polyvinyl alcohol is used to
39 prevent cracking during the sintering process. The viscosity of the nanosilver paste is
40 about 174 Pa-s, the weight percentage of silver in the paste is 83.5%, and the heat
41 absorption temperature is 335.4°C. The Ag nanoparticles were mixed with organics
42 by stirring to form a uniformly nano-Ag paste in the lab before sample manufactur-
43 ing. The die material is Si, and to enhance the interface bonding strength and avoid
44 delamination during the test due to the material difference between Si and nanosilver,
45
the bottom side of the die was sputtered with Ti and Ag, successively. Similarly, in
46
order to enhance the affinity between the copper substrate and the nanosilver solder
47
48 paste, the upper surface of the substrate is covered with an Ag layer.
49
50 9
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
Prior to sintering, the nanosilver paste is placed in the center of the substrate by
1 hand scraping, using a stencil mask with a thickness of 100 µm and a window size of
2 5×5 mm. The Silicon die is placed on the surface of the solder paste and subsequently
3 sintered.
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21 Fig. 4: The manufacturing process of nanosilver sintered joint samples
22
23
24
25
26
27 Table I The properties of nanosilver sintered sample components
28 Sample component Die Bonding layer Substrate
29
Material Silicon [43] Sintered nanosilver Copper [44]
30
Young’s modulus (MPa) 1.31×105 4.0×104 [45] 1.2×105
31 Poisson’s ratio 0.22 0.38 [46] 0.34
32 Density (kg/m3 ) 2330 8600 [42] 8960
33 Length (mm) 5 5 10
34 Width (mm) 5 5 10
35 Thickness (mm) 1 0.1 3
Surface treatment Ti (50 nm),Ag (50 nm) / Ag (50 µm)
36
37
38
39 The samples were manufactured with the pressure-assisted sintering process [47],
40 where the sintering pressure is 1 MPa. The sintering temperature (T s ) is 275°C and
41 300°C, respectively and loaded as the curve plotted in Fig. 4. During the first stage
42 (from room temperature to 150°C), low boiling point organics (e.g., glycol and water)
43 begin to decompose. which reduces the fluidity of the solder paste, thus preventing the
44 die from being buried and shorted by the extruded paste. At the second stage (150°C
45
- 250°C), the organics of larger molecular weight (e.g., polyethylene glycol) gradually
46
begin to decompose and the solder paste weight decreases and remains stable for a
47
48 period of time, enabling further curing of the paste. For the last stage, the polyvinyl
49
50 10
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
Table II Sintering sample list and corresponding sintering condi-
1
tions
2
3 Sintering Sintering Sintering
Sample ID Experiment
4 temperature (°C) time (min) pressure (MPa)
5 275-1 275 15 1 SEM
6 275-2 275 15 1 Shear test
7 275-3 275 15 1 Shear test
300-1 300 15 1 SEM
8 300-2 300 15 1 Shear test
9 300-3 300 15 1 Shear test
10
11
12
13 alcohol decomposes, and the solder paste sintering and curing is completed with the
14 temperature keeping time of T s is 15 min.
15 Following the above process, 6 samples, tabulated in Table II, were prepared for
16 SEM observation and shear test to verify the proposed modeling method in this paper.
17
18
19
20 3.2 SEM observation
21 As shown in Fig. 5, the nanosilver sintered joint observation specimens were fabricated
22 using the resin inlay method, grinded to the center of the specimen, and polished
23 using a diamond polishing compound with a particle size of 0.5 µm to obtain a fine
24 sintered silver cross-section. Using a ZEISS SEM, the microscopic porous structures of
25
the nanosilver sintered joints sintered at T s of 275°C and 300°C degrees are captured
26
and shown in Fig. 6.
27
28
29
30
31
32
33
34
35
36
37
38
39 Fig. 5: SEM observation flow of nanosilver sintered joint samples
40
41
42
43 By comparing the SEM figures, it can be found that when the sintering temperature
44 is increased from 275°C to 300°C, there is the phenomenon of the sintering neck growth,
45
the contact points between adjacent silver particles, inside the sintered joint. And at
46
the same time, the smaller stochastically distributed voids are forced to join resulting
47
48 in increasing volume and decreasing density.
49
50 11
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12 Fig. 6: Microstructure SEM observation results with the scale of 2 µm of nanosilver
13 sintered joints at the sintering temperature of (a)275°C and (b) 300°C, respectively
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Fig. 7: The shear section schematic diagram of the nanosilver sintered joint samples
35
36
37
38
39 3.3 Shear test
40 The shear tests of samples sintered under Ts of 275°C and 300°C was carried out by
41 the Dage Series 4000 Bond tester, where the sample was fixed on a vacuum suction
42 cup and shear force was applied to the silicon die by a wedge-shaped shear tool with
43 a shear height relative to the upper surface of the substrate of 100 µm. The shear
44 tool moved at a rate of 200 µm/s and stopped when the die was removed from the
45
nanosilver sintered joint. The force captured by sensors at that time was termed as
46
the shear force and recorded in Table III, and based on Eq. 3, the shear strength was
47
48 calculated accordingly.
49
50 12
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2 F
τ= (3)
3 A
4
5 where τ in MPa is the shear strength, F in N is the shear force and A is the shear
6 section area which is equal to 25 mm2 .
7 It can be seen from Fig. 7, the schematic diagram of the shear section, the failure
8 phenomenon of die breakage, sintered layer damage and sintered layer peeling from the
9 substrate may occur in the cross section when the nanosilver sintered joint is subjected
10 to the shear damage, where the sintered layer damage takes the dominant percentage,
11 so the shear test developed in this section can be used to evaluate the shear properties
12 of the nanosilver sintered joint. Concluding from the results of shear strength in Table
13 III, the increased sintering temperature of the nanosilver sintered joint can lead to a
14 better capability to against the shear load.
15
16
17 Table III Nanosilver sintered joint samples shear test result
18
Sample ID Sintering Temperature (°C) Shear force (N) Shear strength (MPa)
19
20 275-2 275 390.90 15.64
21 275-3 275 346.56 13.86
300-2 300 873.32 34.93
22 300-3 300 991.78 39.62
23
24
25
26
27 4 Results and discussion
28
29 4.1 SREs reconstruction and FE simulation
30
31 Taking the SEM figures in Fig. 6, which demonstrate the microscopic porous structures
32 of the nanosilver sintered joints as the input to reconstruct the corresponding SREs,
33 where the image size is 60 µm×45 µm. The SRE size is set as 20 µm×20 µm which is
34 relatively small when compared to the sintered joint but large enough to contain the
35 structural information. Three equal images of every sample were cut off sequentially
36 from the centered typical sintering area in its SEM figure and shown in Fig. 8, whose
37 porosities were listed in Table IV.
38 It can be seen that the increased sintering temperature leads to a tighter sintering
39 structure with lower porosity. In addition, the similar porosities of three sintering areas
40 from the same sintering sample represent a homogeneous sintered structure, thus the
41 selected areas can be used to generate the SREs for the corresponding sintered joint.
42 Following the reconstruction method in Section 2.2 with a strict similarity tolerance
43 of 2%, the reconstruction parameters for generating the SRE are figured out with 4
44 round iterations and listed in Table IV, where a represents the autocorrelation param-
45
eter in the horizontal direction and b of vertical direction. The larger autocorrelation
46
parameters indicate the increase in sintering temperature leads to larger microscopic
47
48 voids as a result of voids being squeezed and then merged by the growing sintered
49
50 13
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22 Fig. 8: Selected sintering area from experiment samples sintered at (a)-(c) 275°C and
23 (d)-(f) 300°C
24
25
26
27 Table IV The porosity and reconstruction parameters of sintering area
28 Sintering Average Reconstruction
Sample ID Area ID Porosity
29 Temperature (°C) porosity parameter (×103 µm)
30 (a) 0.2115
31 275-1 275°C (b) 0.2227 0.2172 a=338.24, b=286.76
32 (c) 0.2174
33 (d) 0.1189
34 300-1 300°C (e) 0.1275 0.1243 a=500.00, b=360.29
(f) 0.1265
35
36
37
38 neck, which is consistent with the conclusions in Section 3.2. This phenomenon can
39 also be observed from Fig. 8 and the two reconstructed SRE plotted in Fig. 9.
40 Referring to the method described in Section 2.3.1, the SREs plotted in Fig. 9 were
41 modeled in ANSYS for further simulated calculation. The input material parameters
42 are listed in Table I. As for meshing, the element type is chosen as PLANE182 to
43 simulate the elastic behavior. The maximum element size is 58.8×10−3 µm, which is
44 1 pixel in SRE bitmap.
45
To get the equivalent Young’s modulus and Poisson’s ratio, the tensile simulation
46
as described in Section 2.3.2 was executed, where a 1 MPa uniform load is applied
47
48 to the right boundary and other constraint conditions were tabulated in Table V and
49
50 14
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
depicted in Fig. 10, where the coupling DOF means the coupled nodes have identical
1 motion.
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16 Fig. 9: Reconstructed SRE of experiment sample sintered at (a) 275℃, (b)300℃
17
18
19
20
21
22 Table V SRE finite element simulation bound-
23 ary conditions and loads
24 Location Boundary condition Object Value
25
X = 0 dx = 0 /
26
X = 20 µm Coupling DOF /
27 X = 20 µm Uniform load P Nodes 1 MPa
28 Y = 0 dy = 0 /
29 Y = 20 µm Coupling DOF /
30
31
32
33 The simulation results are shown in Fig. 11, where the displacement ranges are
34 0.117 × 10−2 µm and 0.667 × 10−3 µm, respectively. The displacement differences along
35 the tensile direction prove that stochastically distributed voids lead to the inhomoge-
36 neous displacement distribution which should present a uniform gradient change in a
37 homogeneous material, the larger the size of the voids along the displacement direction,
38 the greater the displacement gradient. Besides, the increased porosity of the sintered
39 joints with lower sintering temperatures aggravates the deformation of corresponding
40 SRE and worsens the elastic tensile resistance.
41 Extracting the deformation of SRE in x-direction and y-direction and according to
42 Eq. 1 and 2, the equivalent Young’s modulus and equivalent Poisson’s ratio of SREs
43 are calculated as Eq. 4 to 7, where the subscripts stand for the sintering tempera-
44 ture. Compared to input Young’s modulus and Poisson’s ratio values, 40 MPa and
45
0.38, respectively, the existence of stochastically distributed voids leads to a drop in
46
the equivalent Young’s modulus and Poisson’s ratio, whose magnitudes enlarge with
47
48 increasing porosity.
49
50 15
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12 Fig. 10: The (a) boundary conditions and (b) load of the SRE finite element simulation
13 model, where the light blue arrows represent the displacement constraints, the green
14 arrows represent the coupling degree of freedom, and the red arrows represent the
15 uniform tensile load, respectively
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32 Fig. 11: Simulated displacement results of SRE sintered at (a) 275°C and (b) 300°C,
33 when subjected to a nodal tensile load of 1 MPa
34
35
36
37
38 P lx275 1MPa × 20µm
39 E275 = = = 17094.02MPa (4)
∆lx275 0.117 × 10−2 µm
40
41
42 P lx300 1MPa × 20µm
43 E300 = = = 29985.01MPa (5)
∆lx300 0.667 × 10−3 µm
44
45
46 ∆ly275 0.377 × 10−3 µm
47 ν275 = − = = 0.3230 (6)
48 ∆lx275 0.117 × 10−2 µm
49
50 16
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12 Fig. 12: The probability density function curves of SRE with sintering temperature,
13 (a) equivalent Young’s modulus, (b) equivalent Poisson’s ratio
14
15
16
17
18
19 ∆ly300 0.245 × 10−3 µm
20 ν300 = − = = 0.3682 (7)
∆lx300 0.667 × 10−3 µm
21
22
23
24 4.2 The Monte Carlo-based FE simulation and parameters
25 fitting
26
27 To scale the effect of the stochastically distributed voids on the dispersion of the
28 nanosilver sinter joint’s properties, as the method proposed in Section 2.3.2, 50 sets
29 of SREs are reconstructed based on the same pair of reconstruction parameters listed
30 in Table IV and imported into the Monte Carlo-based simulation process to get the
31 results of equivalent Young’s modulus and Poisson’s ratio.
32 The probability density function curves of the equivalent constitutive parameters
33 of SREs with sintering temperature are plotted in Fig. 12. Compared to the solid lines
34 representing the T s of 300°C, the dashed lines shift towards the left of the coordinate
35 axis, which means the equivalent Young’s modulus and equivalent Poisson’s ratio of
36 SREs decrease as the increased porosity of nanosilver sintered joint caused by the
37 lower sintering temperature. In addition, the distribution range of these constitutive
38 parameters at lower sintering temperatures is larger. It can be speculated that when
39 the sintering temperature goes down, the microscopic voids are distributed in a more
40 complex pattern due to the increased porosity and reduced average void size, as a
41 consequence, the dispersion of parameter variations improves.
42
With the highest acceptance probability values of 0.9944 and 0.9966, the prob-
43
ability density functions of the equivalent Young’s modulus are fitted into the
44
45 three-parameter Weibull distribution with the stochastic distribution form shown in
46 Eq. 8, and the fitted stochastic distribution for the equivalent Poisson’s ratio is the
47 Normal distribution with the function as Eq. 9. All the fitted parameters are listed in
48 Table VI.
49
50 17
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1 " 
2  β−1 β #
β x−γ x−γ
3 f (x) = exp − (8)
η η η
4
5
6 where η is the scale parameter, β is the shape parameter and γ is the location
7 parameter.
8
9 " 
10 2 #
1 x−µ
11 f (x) = √ exp − √ (9)
12 2πσ 2σ
13
14 where µ is the mean value and σ is the standard deviation.
15
16
17 Table VI The equivalent parameters fitting results
18
19 Sintering temperature (°C) 300 275 300 275
20 Parameter E ν
21 Distribution form Three-parameter Weibull Normal
ν/η 1.2133 2.6106 0.3632 0.3275
22
σ/β 2540 4150 0.008 0.0158
23 γ 26300 14000 / /
24 Acceptance probability 0.9944 0.9966 0.9670 0.9925
25
26
27
28
29 4.3 The multiscale modeling for nanosilver sintered joint
30
31 Referring to the design dimensions of the test samples, a macro scale simulation model
32 of the nanosilver sintered joint sample is built as shown in Fig. 13, where the block
33 from top to bottom is the silicon die, nanosilver sintered joint and copper substrate,
34 respectively. Considering the applied pressure and decomposition of organic matter
35 during the nanosilver sintering process, the modeled thickness of the sintered layer is
36 60 µm, and other dimensions are consistent with the values in Table I.
37
38
39
40
41
42
43
44
45
46
47 Fig. 13: The geometry model for macro finite element simulation
48
49
50 18
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
For this 3D model, the mesh element type is set as SOILID185 to simulate the
1 elastic behavior. Following the meshing requirements proposed in Section 2.5, the
2 element size of the sintered joint in X, Y, Z directions is determined as 20 µm, the
3 same as the SRE’s dimension. The same value is also used as the element size of the
4 silicon die and the copper substrate in X, Y directions to guarantee the continuous
5 nodes at the interfaces.
6 The material properties of the die and substrate and the density of the nanosilver
7
sintered joint are listed in Table I, while Young’s modulus and Poisson’s ratio of each
8
element in the sintered joint are sampled from the fitted distribution in last section.
9
10 Benefiting from this, the sintered joint is capable of characterizing the stochastic influ-
11 ence from micro voids without the need to model them physically. In this model,
12 187500 pairs of Young’s modulus and Poisson’s ratio values are sampled from the
13 Three-parameter Weibull distribution and the Normal distribution, respectively.
14
15 4.4 Nanosilver sintered joint FE simulation and experimental
16 verification
17
18 To verify the applicability of the FE model for nanosilver sintered joint built based
19 on the computational multiscale modeling method proposed in this paper, a FE sim-
20 ulation to get the shear strength as the experiment in Section 3.3 is developed as
21 follows.
22 Referring to the shear test conditions, the lower surface and the two sides along the
23 Y-axis direction of the substrate are set as completely fixed constraints to simulate the
24 fixture offered by a vacuum suction cup. A uniform load is applied on a side surface
25 of the die which is parallel to the X-Z plane with a coordinate of Y = -2.5 µm and the
26 direction is along the positive Y-axis., where the value of the load is calculated from
27
the critical shear force measured in the shear test as the Eq. 10.
28
29
30
31 Pshear = F/Across (10)
32
33 where P shear in MPa is the uniform load used in the simulation, F in N is the shear
34 force measured in the experiment and recorded in Table III, and Across represents the
35 area of surface that the shear load Pshear acting on, whose value is 300 µm2 .
36 In addition, the silicon die is set as a completely rigid body by coupling all degrees
37 of freedom of all its nodes, without considering the fracture of the silicon die during
38 shearing. The simulation boundary constraints and load conditions are shown in Fig.
39 14.
40 The simulation results of the stresses in the Y-Z plane of the nanosilver sintered
41 joint corresponding model to the sample 275-2 and the zoomed bonding layer are
42 shown in Fig. 15. It can be seen that the maximum shear stress in nanosilver sintered
43 joint occurs at the edge of the sintered layer with the greatest deformation during
44 the shear process. For other samples, the stress distribution patterns are analogous.
45
In this process, the shear deformation occurred in Y-Z plane, so with reference to
46
the factors of loading position, loading direction, and the shear strength calculation
47
48 method, the average shear stress response in Y-Z direction of the sintered layer nodes
49
50 19
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12 Fig. 14: The (a) boundary conditions and (b) load of the macro finite element sim-
13 ulation model for nanosilver sintered joint, where the light blue arrows represent the
14 displacement constraints, the green arrows represent the coupling degree of freedom,
15 and the red arrows represent the uniform pressure load, respectively
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33 Fig. 15: The simulation results of the shear stress in the Y-Z plane of (a) the sintered
34 sample 275-2 and (b) corresponding bonding layer
35
36
37
38 can be extracted as the simulated shear strength and compared with the experimental
39 results from Section 3.3 and researches [48] and [49] where the samples are also sintered
40 under 1MPa, 275°C and 300°C, respectively, which is plotted in Fig. 16.
41 It can be seen that with the increase of the sintering temperature, the shear test
42 results and the finite element simulation results based on the computational multiscale
43 modeling method consistently showed the trend of increasing shear strength of the
44 joints; in addition, the simulation results are fairly consistent with the shear test
45
results, although there was a certain error. For the simulation, the error may have come
46
from the relatively small simulation input of Young’s modulus, which was obtained
47
48 from the nanoindentation test on the non-fully dense specimen.
49
50 20
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
Fig. 16: The simulated shear strength of the nanosilver sintered joints versus the
17
18 experimental results
19
20
21
22 Based on the classical double-sphere model [50] and Coble’s tetrahedron model
23 [51] describing the sintering process of material particles, elevating the sintering tem-
24 perature effectively promotes the sintering process and reduces the porosity of the
25 sintered body, creating more sintering necks and enriching the grain boundaries. The
26 presence of grain boundaries at room temperature can hinder the movement of dis-
27 locations, thus increasing the material’s resistance to plastic deformation. Therefore,
28 for nanosilver sintered joints, increasing the sintering temperature can enhance the
29 shear strength of the joints, consistent with the results of finite element simulations
30 conducted based on the multi-scale model proposed in this paper. Furthermore, for
31 nanosilver-sintered materials, strength and stiffness generally have a proportional rela-
32
tionship. Hence, nanosilver sintered joints obtained at higher sintering temperatures
33
34 have greater resistance to elastic deformation, and a higher Young’s modulus, which
35 shows consistent trends with the finite element simulation results of the nanosilver
36 sintered joint SRVE in microscale.
37 The above analysis declares that the simulation of nanosilver sintered joints based
38 on the computational multi-scale modeling method proposed in this study can prop-
39 erly predict the constitutive parameters distribution and the mechanical properties of
40 sintered joints, and the model can achieve more accurate prediction when the accuracy
41 of the input parameters is highly guaranteed.
42
43
44
45 5 Conclusions
46
A computational multiscale modeling method for nanosilver sintered joints with
47
48 stochastically distributed voids is originally proposed in this paper. Through the
49
50 21
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
finite element (FE) simulation on the micro-scale porous structure and the macro-
1 scale sintered samples with experimental verification, the following conclusions can be
2 drawn:
3 (1) A multiscale modeling method for nanosilver sintered joint is developed in 4
4 steps, including reconstruction of the stochastic representative elements (SREs) for
5 micro porous morphology, Monte Carlo-based FE simulation on SREs, distribution
6 fitting for constitutive parameters, and stochastically sampling for macro modeling.
7
This proposed method makes it possible to analyze the influence of the micro voids
8
on the macro properties in a simulation without modeling the micro details in the
9
10 simulation model.
11 (2) On the microscale, an approach to convert the morphology information of SRE
12 into an operational FE model is proposed where the SREs are reconstructed based on
13 the characteristics of the stochastically distributed voids. The tensile simulation on the
14 FE model shows that the stochastically distributed voids lead to the inhomogeneous
15 displacement distribution and the increased porosity of the sintered joints with lower
16 sintering temperature aggravates the deformation of corresponding SRE and worsens
17 the elastic tensile resistance.
18 (3) The distribution of equivalent constitutive parameters is captured from the
19 Monte Carlo-based FE simulation and parameters fitting and used as the link to
20 transfer the effect of the microscopic voids to the macroscopic scale of the sintered
21 joint. Influenced by the stochastic distribution of the voids, the equivalent Young’s
22 modulus shows a trend of three-parameter Weibull distribution and the equivalent
23 Poisson’s ration is Normal distribution. When the sintering temperature goes down,
24 the microscopic voids are distributed in a more complex pattern due to the increased
25
porosity and reduced average void size, which are observed in the experimental SEM
26
images and Monte Carlo-based simulation models, then the dispersion of parameter
27
28 variations improves.
29 (4) The proposed modeling method is used to simulate the shear process of the
30 nanosilver sintered samples and verified by the shear test, where the simulated and
31 experimental results of the shear strength show acceptable agreements, which turns
32 up the applicability of the proposed method. Both the simulation and the experiment
33 show that increasing the sintering temperature can improve the ability of nanosilver
34 sintered joints to resist shear damage, which is beneficial to reduce the possibility
35 of thermal fatigue damage in the solder layer of electronic packaging structures and
36 improve the inherent reliability level of nanosilver sintered materials.
37
38 Acknowledgments. The authors would like to thank Associate Professor Cheng
39 Qian and Associate Professor Hongqiang Zhang for their kind help in method
40 development and experiment, respectively.
41
42 Conflict of Interest. The authors declare that they have no conflict of interest.
43
44
45 Statements & Declarations
46 • Funding
47
This research is carried out as a part of the Center of Digitalized Electronics (CoDE)
48
49
50 22
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
project at Aalborg University funded by the Poul Due Jensen Grundfos Foundation
1 (Grant number 2021-001).
2 • Author Contributions
3 Zhongchao Sun: Conceptualization, Methodology, Software, Formal analysis and
4 investigation, Visualization, Writing - original draft preparation;
5 Wendi Guo: Conceptualization, Methodology, Software, Writing - review and edit-
6 ing;
7
Asger Bjørn Jørgensen: Visualization, Writing - review and editing, Funding
8
acquisition.
9
• Data availability
10
11 The data sets generated and analyzed during the current study are available from
12 the corresponding author upon reasonable request.
13 • Ethics approval
14 This research did not contain any studies involving animal or human participants,
15 nor did it take place on any private or protected areas.
16
17
18
Appendix A APDL code for building SRE FE model
19 The core code run in the ANSYS APDL environment for building SRE FE model is
20 shown below:
21
22 / PREP7
23 ! Define the model size
24 LENGTH =340
25 WIDTH =340
26 TOL =0.5 E -4
27 DOTNUMBER = LENGTH * WIDTH
28 MEDIUMNUMBER =25108
29
! Define the unit length
30
UL =0.0588
31
32 ! Draw the key points
33 FLAG =1
34 i =1
35 * DOWHILE , FLAG
36 COLUMN = MOD (i , LENGTH )
37 * IF , COLUMN , EQ ,0 , THEN
38 COLUMN = LENGTH
39 * ENDIF
40 ROW =( i - COLUMN )/ LENGTH +1
41 K , ,( COLUMN -1)* UL ,( ROW -1)* UL ,0
42 K , , COLUMN * UL ,( ROW -1)* UL ,0
43 K , , COLUMN * UL , ROW * UL ,0
44 K , ,( COLUMN -1)* UL , ROW * UL ,0
45 i = i +1
46 * IF ,i , GT , DOTNUMBER , THEN
47
FLAG =0
48
49
50 23
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
* ENDIF
1 * ENDDO
2 ! Load the . txt file recording the SRE morphology
3 * DIM , MEDIUM , ARRAY , MEDIUMNUMBER ,1 ,1
4 * VREAD , MEDIUM , MEDIUMFile , TXT
5 ( F15 .0)
6 ! Build the areas corresponding to the medium
7
FLAG =1
8
i =1
9
10 * DOWHILE , FLAG
11 p = MEDIUM ( i )
12 A ,4* p -3 ,4* p -2 ,4* p -1 ,4* p
13 i = i +1
14 * IF ,i , GT , MEDIUMNUMBER , THEN
15 FLAG =0
16 * ENDIF
17 * ENDDO
18 ALLSEL , ALL
19 ASEL ,S , AREA , ,1 , MEDIUMNUMBER ,1
20 CM , MEDIUMS , AREA
21 ! Connect all separate medium ( sintered nanosilver )
22 ALLSEL , ALL
23 ASEL ,S , AREA , ,1 , MEDIUMNUMBER ,1 ,1
24 NUMMRG , KP
25
26
27 Appendix B APDL code for SRE FE simulation
28
29 The core code run in the ANSYS APDL environment for Monte Carlo-based FE
30 simulation of SREs is shown below:
31
32 / PREP7
33 ! Define the number of SREs
34 MODELNUMBER =50
35 ! Calculate the equivalent Young ’ s modulus and Poisson ’ s ratio
36 * DO , model ,1 , MODELNUMBER ,1
37 ! Build the SRE FE model
38 ...
39 Code block 1
40 ...
41 ! Solve the model
42 / SOL
43 ANTYPE , STATIC
44 ! Define the fixed boundary
45 ALLSEL , ALL
46 NSEL ,S , LOC ,X ,0 - TOL ,0+ TOL
47
NPLOT
48
49
50 24
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
D , ALL , UX ,0
1 D , ALL , UY ,0
2 ! Coupling the degree of freedom and Define the load
3 ALLSEL , ALL
4 NSEL ,S , LOC ,X , LENGTH * UL - TOL , LENGTH * UL + TOL
5 NPLOT
6 CP , NEXT , UX , ALL
7
CP , NEXT , UY , ALL
8
SF , ALL , PRES , -1
9
10 ALLSEL , ALL
11 SOLVE
12 FINISH
13 ! Extract the results
14 ! Calculate the equivalent Young ’ s modulus
15 ! Calculate the equivalent Poisson ’ s ratio
16 ...
17 Code block 2
18 ...
19 ! Clear current model and prapare for next model
20 / PREP7
21 EDELE , ALL
22 NDELE , ALL
23 ADELE , ALL
24 KPDELE , ALL
25
* ENDDO
26
27
28 References
29
30 [1] Wang, Y., Ding, Y., Yin, Y.: Reliability of Wide Band Gap Power Electronic
31 Semiconductor and Packaging: A Review. Energies 15(18) (2022) https://doi.
32 org/10.3390/en15186670
33
34 [2] Jorgensen, A.B., Munk-Nielsen, S., Uhrenfeldt, C.: Overview of Digital Design
35 and Finite-Element Analysis in Modern Power Electronic Packaging. IEEE Trans-
36 actions on Power Electronics 35(10), 10892–10905 (2020) https://doi.org/10.
37 1109/TPEL.2020.2978584
38
39 [3] Lad, A.A., Hoque, M.J., Christian, S., Zhao, Y., Balda, J.C., King, W.P.,
40 Miljkovic, N.: High power density thermal management of discrete semiconduc-
41 tor packages enabled by additively manufactured hybrid polymer-metal coolers.
42
Applied Thermal Engineering 220(August 2022), 119726 (2023) https://doi.org/
43
10.1016/j.applthermaleng.2022.119726
44
45
[4] Suganuma, K.: Advances in lead-free electronics soldering. Current Opinion in
46
Solid State and Materials Science 5(1), 55–64 (2001) https://doi.org/10.1016/
47
S1359-0286(00)00036-X
48
49
50 25
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
[5] Abtew, M., Selvaduray, G.: Lead-free solders in microelectronics. Materials Sci-
1 ence and Engineering R: Reports 27(5), 95–141 (2000) https://doi.org/10.1016/
2 S0927-796X(00)00010-3
3
4 [6] Zhao, K., Zhao, J., Wei, X., Zhang, X., Deng, C., Yang, Y.: Mechanical
5 properties and microstructure of large-area diamond / silicon bonds formed
6 by pressure-assisted silver sintering for thermal management. Materials Today
7 Communications 34(December 2022), 105230 (2023) https://doi.org/10.1016/j.
8 mtcomm.2022.105230
9
10 [7] Yoon, J.W., Back, J.H.: Metallurgically and mechanically reliable microsilver-
11 sintered joints for automotive power module applications. Journal of Materials
12 Science: Materials in Electronics 33(3), 1724–1737 (2022) https://doi.org/10.
13
1007/s10854-022-07728-6
14
15 [8] Wang, C., Zhang, X., Zhang, Y., Zhao, T., Zhu, P., Sun, R., Nishikawa, H., Xu,
16
L.: Pressureless and low temperature sintering by Ag paste for the high temper-
17
ature die-attachment in power device packaging. In: 2022 IEEE 72nd Electronic
18
19 Components and Technology Conference (ECTC), vol. 2022-May, pp. 2256–2262.
20 IEEE, San Diego, USA (2022). https://doi.org/10.1109/ECTC51906.2022.00356
21
22 [9] Kobayashi, T., Ando, T.: Recent development of joining and conductive materials
23 for electronic components. Materials Transactions 62(8), 1270–1276 (2021) https:
24 //doi.org/10.2320/matertrans.MT-M2021060
25
26 [10] Chen, C., Kim, D., Zhang, Z., Wakasugi, N., Liu, Y., Hsieh, M.C., Zhao, S.,
27 Suetake, A., Suganuma, K.: Interface-Mechanical and Thermal Characteristics
28 of Ag Sinter Joining on Bare DBA Substrate during Aging, Thermal Shock and
29 1200 W/cm2Power Cycling Tests. IEEE Transactions on Power Electronics 37(6),
30 6647–6659 (2022) https://doi.org/10.1109/TPEL.2022.3142286
31
32 [11] Zhou, H., Guo, K., Ma, S., Wang, C., Fan, X., Jia, T., Zhang, Z., Xu, H., Xing, H.,
33 Wang, D., Liu, C.: A triple-layer structure flexible sensor based on nano-sintered
34 silver for power electronics with high temperature resistance and high thermal
35 conductivity. Chemical Engineering Journal 432(December 2021), 134431 (2022)
36 https://doi.org/10.1016/j.cej.2021.134431
37
38 [12] Chen, H., Zhang, J., Zhang, G., Liu, P.: Review of Laser Sintering of Nanosilver
39 Pastes for Die Attachment: Technologies and Trends. In: 2022 23rd International
40 Conference on Electronic Packaging Technology, ICEPT 2022. IEEE, ??? (2022).
41 https://doi.org/10.1109/ICEPT56209.2022.9872618
42
43 [13] Wang, J., Chen, S., Zhang, L., Zhao, X., Duan, F., Chen, H.: Brief Review
44 of Nanosilver Sintering: Manufacturing and Reliability. Journal of Electronic
45 Materials 50(10), 5483–5498 (2021) https://doi.org/10.1007/s11664-021-09078-1
46
47 [14] N’Tsouaglo, K.H., Milhet, X., Colin, J., Signor, L., Nait-Ali, A., Creus, J.,
48
49
50 26
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
Gueguen, M., Gadaud, P., Legros, M.: Time-Resolved Evolution of the 3D
1 Nanoporous Structure of Sintered Ag by X-Ray Nanotomography: Role of the
2 Interface with a Copper Substrate. Advanced Engineering Materials 24(1), 1–13
3 (2022) https://doi.org/10.1002/adem.202100583
4
5 [15] Wakamoto, K., Mochizuki, Y., Otsuka, T., Nakahara, K., Namazu, T.: Tempera-
6 ture dependence on tensile mechanical properties of sintered silver film. Materials
7 13(18) (2020) https://doi.org/10.3390/ma13184061
8
9 [16] Wereszczak, A.a., Vuono, D.J., Wang, H., Ferber, M.K.: Properties of Bulk
10 Sintered Silver As a Function of Porosity, pp. 1–36 (2012)
11
12 [17] Namazu, T.: Mechanical Property Measurement of Micro/Nanoscale Materials for
13 MEMS: A Review. IEEJ Transactions on Electrical and Electronic Engineering,
14 308–324 (2022) https://doi.org/10.1002/tee.23747
15
16 [18] He, G., Yao, Y., Yuting, Y.: Size effect on the fracture of sintered porous nano-
17 silver joints: Experiments and Weibull analysis. Journal of Alloys and Compounds
18 863, 158611 (2021) https://doi.org/10.1016/j.jallcom.2021.158611
19
20 [19] Wang, L., Ding, T., Gu, L., Sun, X.: Study on layer formation behavior of Ag
21 joints sintered with pressureless sintering process. Materials Research Express
22
9(11), 116512 (2022) https://doi.org/10.1088/2053-1591/ac9d84
23
24 [20] Wakamoto, K., Otsuka, T., Nakahara, K., Namazu, T.: Degradation mechanism
25
of pressure-assisted sintered silver by thermal shock test. Energies 14(17), 1–15
26
(2021) https://doi.org/10.3390/en14175532
27
28
[21] Chen, C., Zhang, Z., Kim, D., Sasamura, T., Oda, Y., Hsieh, M.C., Iwaki, A.,
29
Suetake, A., Suganuma, K.: Interface reaction and evolution of micron-sized Ag
30
31 particles paste joining on electroless Ni-/Pd-/Au-finished DBA and DBC sub-
32 strates during extreme thermal shock test. Journal of Alloys and Compounds
33 862, 158596 (2021) https://doi.org/10.1016/j.jallcom.2021.158596
34
35 [22] Yeo, S.M., Yow, H.K., Yeoh, K.H., Ishak, S.H.B.: Vacuum Reflow Process Opti-
36 mization for Solder Void Size Reduction in Semiconductor Packaging Assembly.
37 IEEE Transactions on Components, Packaging and Manufacturing Technology
38 12(8), 1410–1420 (2022) https://doi.org/10.1109/TCPMT.2022.3189995
39
40 [23] Gong, H., Wang, T., Zhu, J., Li, S., Yao, Y.: Compressive experimental analy-
41 sis and constitutive model of sintered nano-silver. Journal of Applied Mechanics
42 90(March), 1–29 (2022) https://doi.org/10.1115/1.4056253
43
44 [24] Xiong, Z., Wang, X., He, M., Benabou, L., Feng, Z.: Investigation on thermal
45 conductivity of silver-based porous materials by finite difference method. Mate-
46 rials Today Communications 33(June), 104897 (2022) https://doi.org/10.1016/j.
47 mtcomm.2022.104897
48
49
50 27
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
[25] Shen, Y.L., Abdo, M.G., Van Rooyen, I.J.: Numerical Study of Effective Thermal
1 Conductivity for Periodic Closed-Cell Porous Media. Transport in Porous Media
2 143(2), 245–269 (2022) https://doi.org/10.1007/s11242-022-01768-6
3
4 [26] Zhao, Z., Zhang, H., Zou, G., Ren, H., Zhuang, W., Liu, L., Zhou, Y.N.: A
5 Predictive Model for Thermal Conductivity of Nano-Ag Sintered Interconnect
6 for a SiC Die. Journal of Electronic Materials 48(5), 2811–2825 (2019) https:
7 //doi.org/10.1007/s11664-019-06984-3
8
9 [27] Renkuan, L., Hul, L., Haiyang, L., Wei, L., Xiao, W., Ran, Y., Renze, Y., Yue,
10 Y.: Analysis on the Reliability Effect of Solder Voids on the Nanosilver Sintered
11 Press-Pack IGBT. In: 2020 4th International Conference on HVDC, HVDC 2020,
12 pp. 20–25 (2020). https://doi.org/10.1109/HVDC50696.2020.9292846
13
14 [28] Qin, F., Hu, Y., Dai, Y., An, T., Chen, P., Gong, Y., Yu, H.: Crack Effect
15 on the Equivalent Thermal Conductivity of Porously Sintered Silver. Jour-
16
nal of Electronic Materials 49(10), 5994–6008 (2020) https://doi.org/10.1007/
17
s11664-020-08325-1
18
19
[29] Letz, S.A., Zhao, D., März, M.: Mesostructural impact on the macroscopic stress
20
state and yield locus of porous polycrystalline silver. Materials and Design 219,
21
22 110785 (2022) https://doi.org/10.1016/j.matdes.2022.110785
23
[30] Suzuki, T., Terasaki, T., Kawana, Y., Ishikawa, D., Nishimura, M., Nakako, H.,
24
25 Kurafuchi, K.: Effect of manufacturing process on micro-deformation behavior of
26 sintered-silver die-attach material. IEEE Transactions on Device and Materials
27 Reliability 16(4), 588–596 (2016) https://doi.org/10.1109/TDMR.2016.2614510
28
29 [31] Schaal, M., Klingler, M., Wunderle, B.: Silver Sintering in Power Electronics: The
30 State of the Art in Material Characterization and Reliability Testing. In: 2018 7th
31 Electronic System-Integration Technology Conference (ESTC), pp. 1–18. IEEE,
32 ??? (2018). https://doi.org/10.1109/ESTC.2018.8546498
33
34 [32] Sun, Z., Wang, Z., Qian, C., Ren, Y., Feng, Q., Yang, D., Sun, B.: Characterization
35 of stochastically distributed voids in sintered nano-silver joints. In: 2019 20th
36 International Conference on Thermal, Mechanical and Multi-Physics Simulation
37 and Experiments in Microelectronics and Microsystems, EuroSimE 2019 (2019).
38 https://doi.org/10.1109/EuroSimE.2019.8724592
39
40 [33] Qian, C., Sun, Z., Fan, J., Ren, Y., Sun, B., Feng, Q., Yang, D., Wang, Z.: Char-
41 acterization and reconstruction for stochastically distributed void morphology in
42 nano-silver sintered joints. Materials and Design 196 (2020) https://doi.org/10.
43 1016/j.matdes.2020.109079
44
45 [34] Zou, Y.S., Gan, C.L., Chung, M.H., Takiar, H.: A review of interconnect materials
46 used in emerging memory device packaging: first- and second-level interconnect
47
48
49
50 28
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
materials. Journal of Materials Science: Materials in Electronics 32(23), 27133–
1 27147 (2021) https://doi.org/10.1007/s10854-021-07105-9
2
3 [35] Chen, C., Suganuma, K.: Microstructure and mechanical properties of sintered
4 Ag particles with flake and spherical shape from nano to micro size. Materials
5 and Design 162, 311–321 (2019) https://doi.org/10.1016/j.matdes.2018.11.062
6
7 [36] Li, S., Mao, Y., Liu, W., Hou, S.: A highly efficient multi-scale approach of
8 locally refined nonlinear analysis for large composite structures. Composite Struc-
9 tures 306(June 2022), 116578 (2023) https://doi.org/10.1016/j.compstruct.2022.
10 116578
11
12 [37] Alshahrani, A., Kulasegaram, S., Kundu, A.: Elastic modulus of self-compacting
13 fibre reinforced concrete : Experimental approach and multi-scale simulation.
14 Case Studies in Construction Materials 18(2022), 01723 (2023) https://doi.org/
15 10.1016/j.cscm.2022.e01723
16
17 [38] Hamed Mashhadzadeh, A., Fereidoon, A., Ghorbanzadeh Ahangari, M.: Com-
18 bining density functional theory-finite element multi-scale method to predict
19
mechanical properties of polypropylene/graphene nanocomposites: Experimental
20
study. Materials Chemistry and Physics 201, 214–223 (2017) https://doi.org/10.
21
22 1016/j.matchemphys.2017.08.042
23
[39] Costa, A.P.O., Seabra, M.R.R., Santos, A.D., Ribeiro, L.M.M., Cesar de
24
25 Sa, J.M.A.: Experimental and Numerical Multiscale characterization of a
26 Super Duplex Stainless Steel 25Cr-7Ni-Mo-N. Materials Today Communications
27 33(October), 104903 (2022) https://doi.org/10.1016/j.mtcomm.2022.104903
28
29 [40] Stefaniuk, D., Niewiadomski, P., Musiał, M., Łydżba, D.: Elastic properties of self-
30 compacting concrete modified with nanoparticles: Multiscale approach. Archives
31 of Civil and Mechanical Engineering 19(4), 1150–1162 (2019) https://doi.org/10.
32 1016/j.acme.2019.06.006
33
34 [41] Aduloju, S.C., Truster, T.J.: On topology-based cohesive interface element
35 insertion along periodic boundary surfaces. Engineering Fracture Mechanics
36 205(October 2018), 10–13 (2019) https://doi.org/10.1016/j.engfracmech.2018.10.
37 037
38
39 [42] Zhang, H., Wang, W., Bai, H., Zou, G., Liu, L., Peng, P., Guo, W.: Microstructural
40 and mechanical evolution of silver sintering die attach for SiC power devices
41 during high temperature applications. Journal of Alloys and Compounds 774,
42 487–494 (2019) https://doi.org/10.1016/j.jallcom.2018.10.067
43
44 [43] Tan, L., Liu, P., She, C., Xu, P., Yan, L., Quan, H.: Heat Dissipation Char-
45 acteristics of IGBT Module Based on Flow-Solid Coupling. Micromachines
46 (2022)
47
48
49
50 29
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
[44] Jorgensen, A.B., Cheng, T.H., Hopkins, D., Beczkowski, S., Uhrenfeldt, C., Munk-
1 Nielsen, S.: Thermal characteristics and simulation of an integrated GaN eHEMT
2 power module. In: 2019 21st European Conference on Power Electronics and
3 Applications. Institute of Electrical and Electronics Engineers Inc., ??? (2019).
4 https://doi.org/10.23919/EPE.2019.8915012
5
6 [45] Long, X., Tang, W., Xia, W., Wu, Y., Ren, L., Yao, Y.: Porosity and Young’s
7 modulus of pressure-less sintered silver nanoparticles. In: 2017 IEEE 19th Elec-
8 tronics Packaging Technology Conference, EPTC 2017, vol. 2018-Febru, pp. 1–8
9 (2018). https://doi.org/10.1109/EPTC.2017.8277577
10
11 [46] Yang, H., Wu, J.: Prediction of mechanical properties and fatigue life of nano
12 silver paste in chip interconnection. Research Square PREPRINT (2020)
13
14 [47] Zhang, H.Q., Bai, H.L., Jia, Q., Guo, W., Liu, L., Zou, G.S.: High Electrical
15 and Thermal Conductivity of Nano-Ag Paste for Power Electronic Applica-
16
tions. Acta Metallurgica Sinica 33(11), 1543–1555 (2020) https://doi.org/10.
17
1007/s40195-020-01083-3
18
19
[48] Lei, T.G., Calata, J.N., Lu, G.Q., Chen, X., Luo, S.: Low-temperature sinter-
20
ing of nanoscale silver paste for attaching large-area (>100 mm2) chips. IEEE
21
22 Transactions on Components and Packaging Technologies 33(1), 98–104 (2010)
23 https://doi.org/10.1109/TCAPT.2009.2021256
24
25 [49] Ide, E., Angata, S., Hirose, A., Kobayashi, K.F.: Metal-metal bonding process
26 using Ag metallo-organic nanoparticles. Acta Materialia 53(8), 2385–2393 (2005)
27 https://doi.org/10.1016/j.actamat.2005.01.047
28
29 [50] Kumar, V.: Simulations and Modeling of Unequal Sized Particles Sintering. PhD
30 thesis, The University of Utah (2011)
31
32 [51] Hsueh, H., Evans, A.G., Laboratorv, L.B., Engineering, M., Coble, R.L.: Develop-
33 ment during final intermediate stage sintering I. pore-grain boundary separation.
34 Acta Metallurgica 30, 1269–1279 (1982)
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50 30
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
Marked Manuscript Click here to access/download;Marked
Manuscript;MarkeManuscript.tex

A computational multiscale modeling method


for nanosilver sintered joints with stochastically
distributed voids
Zhongchao Sun1 , Wendi Guo1*, Asger Bjørn Jørgensen1
1* AAU Energy, Aalborg University, Pontoppidanstræde 101, Aalborg,
9220,Denmark.

*Corresponding author(s). E-mail(s): wg@energy.aau.dk;


Contributing authors: zs@energy.aau.dk; abj@energy.aau.dk;

High power density is required in wide band gap power semiconductor packaging,
which has led to the popularity of sintered nanosilver as an interconnecting mate-
rial. However, affected by stochastically distributed voids in its microstructure, this
material practically exhibits instability leading to reduced reliability. In this paper,
a computational multiscale modeling method is originally proposed to simulate the
influence of micro voids on macro properties, providing an efficient tool to analyze the
aforementioned problem. At the micro-scale, the three-parameter Weibull distribu-
tion of the equivalent Young’s modulus and the Normal distribution of the equivalent
Poisson’s ratio are captured by Monte Carlo-based finite element simulation on the
reconstructed stochastic representative elements, where the density and distribution
morphology of micro voids are taken into consideration. At the macro-scale, the effect
of the microscopic voids is transferred through a random sampling process to construct
the multiscale model. The effectiveness and validity of the proposed method are ver-
ified through experimental case studies involving the modeling of nanosilver sintered
joints sintered at temperatures of 275°C and 300°C. Besides, the effects of the sinter-
ing temperature on the dispersion of the micro voids, the distribution fluctuation of
the constitutive parameters, and the mechanical properties are also discussed based
on numerical and experimental results.

1
Graphical Abstract

2
1 Introduction
Benefiting from high breakdown voltages, high operating temperatures, and high
switching frequencies, wide band gap (WBG) semiconductors such as silicon carbide
(SiC) and gallium nitride (GaN) are dominating the industry of electronics packaging
to ensure high power density [1–3]. To release the potential of WBG semiconductors
adequately and reliably, traditional Pb and lead-free solders with lower melting tem-
peratures [4, 5] are being substituted by new interconnecting materials. Considering
the prerequisites for bonding power chips concerning mechanical connection, electri-
cal transportation, and heat dissipation, low temperature [6, 7] sintering nanosilver
(Ag-NP) emerges as a compelling option within various new attachment technologies
[8, 9], featuring a high operating temperature of up to 961.8°C [10] and low electrical
and thermal resistance of 429 W·m−1 K−1 and 1.6 µΩ·cm [11], respectively.
During the sintering process, silver nanoparticles are driven by the surface energy
and defective energy of the sintering paste to coalesce into a bulk solder, where
mechanical pressure and an external heat source are commonly involved to obtain a
denser structure. However, the applied pressure and excessive temperature may cre-
ate defects and cracks on the chip and substrate [12, 13], so they are limited to some
extent. Due to the inherent nature of the sintering mechanism, stochastically dis-
tributed voids inevitably exist in the microstructure of nanoparticle-sintered joints,
subsequently affecting the macro-level mechanical [14], thermal [15], and electrical
properties [16]. By experimental methods [17], the micro voids observed by scanning
electron microscope (SEM) with larger porosity were found to lead to a decrease in
the tensile strength and bonding strength in nanosilver sintered joints of larger joint
cross-sections [18] and inadequate thickness [19]. In the thermal and electrical fields, a
35% increase in porosity will result in approximate 80% and 25% decreases in thermal
conductivity and electrical conductivity of nanosilver sintered joints, respectively [16].
Triggering the stress singularity point [20], the micro voids are always related to the
crack initiation and propagation, then leading to reliability issues [21, 22]. To model
the relationship between the voids and the macro properties, the material functions-
based mathematical methods [23] and the model-based numerical simulation methods
[24–26] are performed. A multiphysics coupling model [27] of a single-chip 3.3kV/50A
nanosilver sintered press-pack IGBT simulated the impacts of the void distribution,
number, and ratio on the electrical-thermal stress, junction temperature, on-state
voltage, and the chip current distribution, while the voids are regarded as perforated
circles with a size of 0.05 mm rather than the actual morphology. The numerical model
proposed by Fei et al. [28] showed the equivalent thermal conductivity of porously sin-
tered silver is affected by the void shape of the circle, ellipse, and rectangle. Another
research [29] used Focused-Ion-Beam tomography to get the real sintering structures
in the silver film and analyzed their influence on the macroscopic stress state and yield
locus of sintered silver by formulating a thermodynamic consistent continuum model,
but the dispersion of the stochastically distributed voids is not included.
To take a more accurate analysis of nanosilver sintered joints considering the effects
of the porous microstructure, the real morphology has to be taken into account rather
than relying on simplifications [30, 31]. At the same time, the dispersion of the voids
needs to be considered to illustrate the property fluctuation under the same sintering

3
process [32]. Facing the problem, the morphology of the stochastically distributed
voids is randomly reconstructed by keeping the same dispersion characteristics in our
previous research [33], which can be used as the modeling input for the finite element
simulation but cannot be used directly. Although in power electronic packaging, the
bonding layers usually have small geometrical dimensions (at the millimeter level) [34],
the nanosilver sintered joint has a 5000-times geometric difference compared to the
microstructure with stochastically distributed voids (at the micron meter level) [35].
If modeling the power module containing microscopic voids directly for simulation
calculations, it will bring an enormous number of meshes and computational resources,
despite the improved simulation accuracy.
As one popular indirect modeling method, the multi-scale theory [36] is used
to analyze the properties of materials such as concrete [37], graphene-reinforced
polypropylene (PP /graphene) nanocomposites [38], and duplex stainless steel materi-
als [39] at the macroscopic scales by using the representative volume elements (RVEs)
to maintain the disturbance from the micro-scale and without modeling all geometric
details [40], which all have good agreements with the available test data. Given that
nanosilver sintered joints exhibit a similar void morphology and multi-scale feature as
the above materials, it is expected that such a theory can also effectively model their
properties. However, the RVE only works on periodic elements [41] while the real micro
morphologies in nanosilver sintered joints are random, and the properties of the micro
porous structure will behave as a stochastic distribution, accordingly. Therefore, the
universal multiscale modeling pipeline does not work on nanosilver sintered joints. To
solve the problem and build the multiscale model of nanosilver sintered joints which
is crucial to analyze the relationship between the micro voids and macro properties,
a computational method including Monte Carlo simulation and sampling is originally
proposed and verified to construct the multiscale model for nanosilver sintered joints
with stochastically distributed voids in this paper.
The remaining parts of this paper are organized as follows: Section 2 presents
detailed information on the proposed approach to model nanosilver sintered joints
from the microscale to the macroscale by the computational method. Section 3 intro-
duces the manufacturing process of samples and the experimental results of the SEM
observation and shear tests. Section 4 provides the application and validation scenarios
for the proposed approach and analyzes the influence of the stochastically distributed
voids, respectively. In the end, the major conclusions from this study are drawn in
Section 5.

2 Methodology
When a nanosilver sintered joint with stochastically distributed microscopic voids
is loaded, voids are not subjected to the load directly, but voids’ density, size and
distribution morphology affect the generation of internal local pressure and thus the
macroscopic response of the sintered joint. In the case of mechanical loads, this effect is
reflected as a change in the macroscopic constitutive relationship of the sintered joint
which will show a certain statistical distribution due to the statistical characteristics
of the arbitrarily distributed microscopic voids.

4
2.1 Overall methodology flow

Fig. 1: The flow diagram of the proposed multiscale modeling method

The whole procedure of building a finite element analysis (FEA) model of nanosil-
ver sintered joints considering the effect of randomly distributed microscopic voids
based on a computational multiscale approach is shown in Fig. 1. This proposed
procedure includes the following steps:
(1) Based on the method of characterizing and reconstructing the random void mor-
phology of the nanosilver sintered joints, which was previously invented and published
in [33] and presented in Section 2.2, reconstruct the stochastic representative elements
(SREs) that are able to characterize the microscopic void distribution features of the
nanosilver sintered joints.
(2) Taking in the SREs reconstructed above, establish corresponding FEA models
according to the method in Section 2.3 and execute the Monte Carlo finite element
simulations on the micro-scale to obtain the stochastic constitutive model responses
from different microscopic void morphologies of the same sintered joint.
(3) Count the probability distribution of the constitutive parameters calculated by
simulation in the previous step, find the most matched statistical model, and analyze
the distribution pattern using the method developed in Section 2.4.
(4) Model the macroscopic nanosilver sintered joint in the sequence of 3D mod-
eling, meshing, material definition, and boundary conditions application, where the
constitutive parameters of each element are sampled from the fitted models in step
(3). The detailed procedure is referred to the instructions in Section 2.5. Then the
model could be solved in FEA with an ability to consider the influence of microscopic
voids on the macroscopic properties of the nanosilver sintered joint without increasing
the geometric complexity.

5
2.2 Reconstruction of voids
The flow diagram of reconstructing SREs of the microstructure of nanosilver sintered
joints is shown in Fig. 2.
(1) Obtain the SEM images recording the microstructure morphology of the
nanosilver sintered joints clearly, separate and distinguish the stochastically dis-
tributed voids from the sintered nanosilver using OTSU algorithm-based threshold
segmentation image processing methods. Extract the porosity and the characteriza-
tion parameters matrix, which includes the void area (A), distance (l ’) and deflection
angle (α) from the geometric center, and the void aspect ratio (δ), noted as P .
(2) Based on the random medium (RM) theory and the porosity-based threshold
segmentation algorithm, generate a considerable number of void morphology samples
with discrete autocorrelation function parameter inputs, and extract their characteri-
zation parameters matrices Qi as step (1) to form a void morphology database, denoted
as Q.
(3) Compare the similarity between matrix P and Qi by using the Jensen-
Shannon (JS) divergence algorithm, find the characterization parameters matrix Qmax
with the highest matching degree, and note its input autocorrelation parameters of
corresponding void morphology as pending reconstruction parameters.
(4) In the similarity calculation, if the similarity is less than the specified tolerance,
the input parameters are refined with reference to pending reconstruction parameters,
and steps (2-3) are repeated until the specified tolerance is satisfied, and the final
reconstructed parameters are obtained.
(5) Following the same method in step (2), establish void morphology samples
with the same void distribution characteristics as the input nanosilver sintered joint
microstructure by using the final reconstructed parameters. Since the effect of the
stochastic parameter used in RM theory, the samples with the same modeling param-
eters show different morphology features, which exhibit statistical randomness. Hence,
the samples here are noted as the SREs of the nanosilver sinter joints.

Fig. 2: Stochastic representative elements reconstruction flow diagram

6
2.3 Monte Carlo-based Finite element simulation
2.3.1 FE modeling of SRE
The SRE created in Section 2.2 stores the microstructure morphology information of
the nanosilver sintered joint in the form of a binarized matrix, and each matrix cell
represents a pixel in the bitmap, where 0 represents the pixel for the void and 1 repre-
sents the medium. In the geometric modeling stage in the FE simulation environment,
a medium pixel in the SRE can be equated to a two-dimensional plane element.
Based on this principle, the proposed method uses an intermediate file to record
the pixelated structure characteristics of the SRE and nondestructively converts the
SRE from a numerical model to a geometric model. The specific process is illustrated
in Fig. 3 and introduced as follows. The core code run in the Ansys Parametric Design
Language (APDL) environment for SRE FE modeling is provided in the Appendix A.
(1) Extract the morphology features of SRE in the format of ID, row, column,
and feature value, marked as N, I, J, and F, respectively, and export it as a .txt file,
which can be recognized by ANSYS Mechanical easily. In this intermediate file for
SRE modeling, the ID is used to record the SRE size, the row and column are used to
locate the pixel points, and the feature value is used to distinguish whether the pixel
characterizes a void or a medium.
(2) Import the .txt file into ANSYS Mechanical and calculate the actual size of
the SRE in both X and Y directions, which is the product of the maximum values of
the row and column and the plotting scale (the actual size of 1 pixel) of the bitmap.
Following the size limit, draw the key points used to describe all the plane elements
in ANSYS geometry space, where the interval of the key points is the length of the
plotting scale.
(3) Read the pixel feature values in the .txt file, identify the medium compo-
nents, and connect the corresponding key points according to their locations to form
2D planes. Connect all two-dimensional planes representing the sintered nanosilver
material by the Boolean operation to build the SRE geometry model.
(4) In the meshing stage, the plotting scale is set as the quadrangular element
size to guarantee its success regardless of the morphological complexity induced by
stochastically distributed voids. Then, the FE model of SRE is prepared for further
simulation.

Fig. 3: The flow to convert the stochastic representative element into finite element
model

7
2.3.2 Monte Carlo-based FE simulation
During the elastic deformation phase, the effect of microscopic stochastically dis-
tributed voids of the nanosilver sintered joint will be reflected as the variation of the
equivalent Young’s modulus and Poisson’s ratio of the SRE by using Hooke’s law to
describe the constitutive relationship of the nanosilver sintered joint.
With the FE simulation, the tensile deformation in the tensile direction and the
transverse deformation in the vertical direction of the SRE subjected to the uniform
tensile load can be extracted, and according to Eq. 1 and 2, the equivalent Young’s
modulus and Poisson’s ratio can be calculated respectively.

P lx
E= (1)
∆lx

where the X-direction is regarded as the tensile direction, E in MPa is the equiva-
lent Young’s modulus of SRE, P in MPa represents the uniform tensile load, lx in µm
is the size of SRE along the tensile direction and ∆l x in µm is the tensile deformation.

εy ∆ly · lx
ν=− =− (2)
εx ∆lx · ly

where ν is the equivalent Poisson’s ratio of SRE, εx and εy are the tensile strain
and transverse strain, l y in µm is the size of SRE along the transverse direction and
∆l y in µm is the transverse deformation.
In order to analyze the effect of the randomness of the microscopic void distribution
on the constitutive relationship of nanosilver sintered joint, the Monte Carlo-based FE
simulation was conducted. By coding an APDL program displayed in the Appendix
B, the different SRE of the same nanosilver sintered joint reconstructed in Section 2.2
are repeatedly imported to update the FE model and simulated to obtain correspond-
ing the equivalent Young’s modulus and equivalent Poisson’s ratio, according to the
method proposed in this Section.

2.4 Parameter fitting


From the Monte Carlo-based FE simulation responses, different SREs of the same
nanosilver sintered joint will result in different equivalent Young’s modulus and Pois-
son’s ratio observations, which come from the same stochastic distribution, but are
independent of each other.
To characterize the microscopic constitutive parameters distribution of the nanosil-
ver sintered joint considering the randomness of the microscopic void distribution, the
proposed method first calculates the frequency distribution of the discrete constitu-
tion parameters observations obtained from the Monte Carlo-based FE simulation in
Section 2.3 and plots it in the form of a continuous probability density function, and
then takes the commonly used statistical distributions including the normal, lognor-
mal, exponential, Weibull and three-parameter Weibull distributions to fit that. The

8
distribution form with the highest acceptance probability is selected as the stochas-
tic distribution of the constitutive parameters, and the distribution parameters are
determined at the same time.

2.5 Sampling and macro modeling


Considering the size difference between the micro voids and the macro sintered joints
and to reduce the computational burden, rather than modeling all the voids in the
joints, the proposed macroscopic simulation modeling method based on Monte Carlo-
based FE simulation and parameters sampling is developed as below:
In the nanosilver sintered joint FE simulation, a macroscopic geometric model
matching the actual size of the simulation object is established first, and the element
size of the joint is controlled to be the same as that of the SRE in the meshing stage.
Considering the propagation of the influence of the microscopic voids on the consti-
tutive relationship of the nanosilver sintered joint along the scale growth, the physical
parameters of each sintered joint element should be consistent with the microstruc-
ture in the macroscopic FE simulation modeling, and considering the randomness
of the microscopic void distribution, the constitutive parameters of each element of
the sintered joint will be arranged as the random sampling samples of the stochastic
distribution of the equivalent Young’s modulus and the equivalent Poisson’s ratio.
Based on the obtained stochastic distribution forms in Section 2.4, the samples
with the same number of elements as the nanosilver sintered joint are generated by a
random number-based direct sampling method, and the samples are assigned to each
element of the nanosilver sintered joint using the APDL program. With this, the mul-
tiscale modeling for nanosilver sintered joints considering the influence of microscopic
stochastically distributed voids on macroscopic properties is completed.

3 Experiments
3.1 Sample manufacture
The manufacturing process of the nanosilver sintered joint samples with typical "sand-
wich" structure is shown in Fig. 4, and the sample components are shown in Table
I. The nanosilver paste [42] consists of spherical and quasi-spherical Ag nanoparti-
cles with an average particle size of 80 nm and 800 nm and organic matters, where
polyethylene glycol is used to adjust the viscosity and polyvinyl alcohol is used to
prevent cracking during the sintering process. The viscosity of the nanosilver paste is
about 174 Pa-s, the weight percentage of silver in the paste is 83.5%, and the heat
absorption temperature is 335.4°C. The Ag nanoparticles were mixed with organics
by stirring to form a uniformly nano-Ag paste in the lab before sample manufactur-
ing. The die material is Si, and to enhance the interface bonding strength and avoid
delamination during the test due to the material difference between Si and nanosilver,
the bottom side of the die was sputtered with Ti and Ag, successively. Similarly, in
order to enhance the affinity between the copper substrate and the nanosilver solder
paste, the upper surface of the substrate is covered with an Ag layer.

9
Prior to sintering, the nanosilver paste is placed in the center of the substrate by
hand scraping, using a stencil mask with a thickness of 100 µm and a window size of
5×5 mm. The Silicon die is placed on the surface of the solder paste and subsequently
sintered.

Fig. 4: The manufacturing process of nanosilver sintered joint samples

Table I The properties of nanosilver sintered sample components


Sample component Die Bonding layer Substrate
Material Silicon [43] Sintered nanosilver Copper [44]
Young’s modulus (MPa) 1.31×105 4.0×104 [45] 1.2×105
Poisson’s ratio 0.22 0.38 [46] 0.34
Density (kg/m3 ) 2330 8600 [42] 8960
Length (mm) 5 5 10
Width (mm) 5 5 10
Thickness (mm) 1 0.1 3
Surface treatment Ti (50 nm),Ag (50 nm) / Ag (50 µm)

The samples were manufactured with the pressure-assisted sintering process [47],
where the sintering pressure is 1 MPa. The sintering temperature (T s ) is 275°C and
300°C, respectively and loaded as the curve plotted in Fig. 4. During the first stage
(from room temperature to 150°C), low boiling point organics (e.g., glycol and water)
begin to decompose. which reduces the fluidity of the solder paste, thus preventing the
die from being buried and shorted by the extruded paste. At the second stage (150°C
- 250°C), the organics of larger molecular weight (e.g., polyethylene glycol) gradually
begin to decompose and the solder paste weight decreases and remains stable for a
period of time, enabling further curing of the paste. For the last stage, the polyvinyl

10
Table II Sintering sample list and corresponding sintering condi-
tions
Sintering Sintering Sintering
Sample ID Experiment
temperature (°C) time (min) pressure (MPa)
275-1 275 15 1 SEM
275-2 275 15 1 Shear test
275-3 275 15 1 Shear test
300-1 300 15 1 SEM
300-2 300 15 1 Shear test
300-3 300 15 1 Shear test

alcohol decomposes, and the solder paste sintering and curing is completed with the
temperature keeping time of T s is 15 min.
Following the above process, 6 samples, tabulated in Table II, were prepared for
SEM observation and shear test to verify the proposed modeling method in this paper.

3.2 SEM observation


As shown in Fig. 5, the nanosilver sintered joint observation specimens were fabricated
using the resin inlay method, grinded to the center of the specimen, and polished
using a diamond polishing compound with a particle size of 0.5 µm to obtain a fine
sintered silver cross-section. Using a ZEISS SEM, the microscopic porous structures of
the nanosilver sintered joints sintered at T s of 275°C and 300°C degrees are captured
and shown in Fig. 6.

Fig. 5: SEM observation flow of nanosilver sintered joint samples

By comparing the SEM figures, it can be found that when the sintering temperature
is increased from 275°C to 300°C, there is the phenomenon of the sintering neck growth,
the contact points between adjacent silver particles, inside the sintered joint. And at
the same time, the smaller stochastically distributed voids are forced to join resulting
in increasing volume and decreasing density.

11
Fig. 6: Microstructure SEM observation results with the scale of 2 µm of nanosilver
sintered joints at the sintering temperature of (a)275°C and (b) 300°C, respectively

Fig. 7: The shear section schematic diagram of the nanosilver sintered joint samples

3.3 Shear test


The shear tests of samples sintered under Ts of 275°C and 300°C was carried out by
the Dage Series 4000 Bond tester, where the sample was fixed on a vacuum suction
cup and shear force was applied to the silicon die by a wedge-shaped shear tool with
a shear height relative to the upper surface of the substrate of 100 µm. The shear
tool moved at a rate of 200 µm/s and stopped when the die was removed from the
nanosilver sintered joint. The force captured by sensors at that time was termed as
the shear force and recorded in Table III, and based on Eq. 3, the shear strength was
calculated accordingly.

12
F
τ= (3)
A

where τ in MPa is the shear strength, F in N is the shear force and A is the shear
section area which is equal to 25 mm2 .
It can be seen from Fig. 7, the schematic diagram of the shear section, the failure
phenomenon of die breakage, sintered layer damage and sintered layer peeling from the
substrate may occur in the cross section when the nanosilver sintered joint is subjected
to the shear damage, where the sintered layer damage takes the dominant percentage,
so the shear test developed in this section can be used to evaluate the shear properties
of the nanosilver sintered joint. Concluding from the results of shear strength in Table
III, the increased sintering temperature of the nanosilver sintered joint can lead to a
better capability to against the shear load.

Table III Nanosilver sintered joint samples shear test result


Sample ID Sintering Temperature (°C) Shear force (N) Shear strength (MPa)
275-2 275 390.90 15.64
275-3 275 346.56 13.86
300-2 300 873.32 34.93
300-3 300 991.78 39.62

4 Results and discussion


4.1 SREs reconstruction and FE simulation
Taking the SEM figures in Fig. 6, which demonstrate the microscopic porous structures
of the nanosilver sintered joints as the input to reconstruct the corresponding SREs,
where the image size is 60 µm×45 µm. The SRE size is set as 20 µm×20 µm which is
relatively small when compared to the sintered joint but large enough to contain the
structural information. Three equal images of every sample were cut off sequentially
from the centered typical sintering area in its SEM figure and shown in Fig. 8, whose
porosities were listed in Table IV.
It can be seen that the increased sintering temperature leads to a tighter sintering
structure with lower porosity. In addition, the similar porosities of three sintering areas
from the same sintering sample represent a homogeneous sintered structure, thus the
selected areas can be used to generate the SREs for the corresponding sintered joint.
Following the reconstruction method in Section 2.2 with a strict similarity tolerance
of 2%, the reconstruction parameters for generating the SRE are figured out with 4
round iterations and listed in Table IV, where a represents the autocorrelation param-
eter in the horizontal direction and b of vertical direction. The larger autocorrelation
parameters indicate the increase in sintering temperature leads to larger microscopic

13
Fig. 8: Selected sintering area from experiment samples sintered at (a)-(c) 275°C and
(d)-(f) 300°C

Table IV The porosity and reconstruction parameters of sintering area


Sintering Average Reconstruction
Sample ID Area ID Porosity
Temperature (°C) porosity parameter (×103 µm)
(a) 0.2115
275-1 275°C (b) 0.2227 0.2172 a=338.24, b=286.76
(c) 0.2174
(d) 0.1189
300-1 300°C (e) 0.1275 0.1243 a=500.00, b=360.29
(f) 0.1265

voids as a result of voids being squeezed and then merged by the growing sintered
neck, which is consistent with the conclusions in Section 3.2. This phenomenon can
also be observed from Fig. 8 and the two reconstructed SRE plotted in Fig. 9.
Referring to the method described in Section 2.3.1, the SREs plotted in Fig. 9 were
modeled in ANSYS for further simulated calculation. The input material parameters
are listed in Table I. As for meshing, the element type is chosen as PLANE182 to
simulate the elastic behavior. The maximum element size is 58.8×10−3 µm, which is
1 pixel in SRE bitmap.
To get the equivalent Young’s modulus and Poisson’s ratio, the tensile simulation
as described in Section 2.3.2 was executed, where a 1 MPa uniform load is applied

14
to the right boundary and other constraint conditions were tabulated in Table V and
depicted in Fig. 10, where the coupling DOF means the coupled nodes have identical
motion.

Fig. 9: Reconstructed SRE of experiment sample sintered at (a) 275℃, (b)300℃

Table V SRE finite element simulation bound-


ary conditions and loads
Location Boundary condition Object Value
X = 0 dx = 0 /
X = 20 µm Coupling DOF /
X = 20 µm Uniform load P Nodes 1 MPa
Y = 0 dy = 0 /
Y = 20 µm Coupling DOF /

The simulation results are shown in Fig. 11, where the displacement ranges are
0.117 × 10−2 µm and 0.667 × 10−3 µm, respectively. The displacement differences along
the tensile direction prove that stochastically distributed voids lead to the inhomoge-
neous displacement distribution which should present a uniform gradient change in a
homogeneous material, the larger the size of the voids along the displacement direction,
the greater the displacement gradient. Besides, the increased porosity of the sintered
joints with lower sintering temperatures aggravates the deformation of corresponding
SRE and worsens the elastic tensile resistance.
Extracting the deformation of SRE in x-direction and y-direction and according to
Eq. 1 and 2, the equivalent Young’s modulus and equivalent Poisson’s ratio of SREs
are calculated as Eq. 4 to 7, where the subscripts stand for the sintering tempera-
ture. Compared to input Young’s modulus and Poisson’s ratio values, 40 MPa and
0.38, respectively, the existence of stochastically distributed voids leads to a drop in
the equivalent Young’s modulus and Poisson’s ratio, whose magnitudes enlarge with
increasing porosity.

15
Fig. 10: The (a) boundary conditions and (b) load of the SRE finite element simulation
model, where the light blue arrows represent the displacement constraints, the green
arrows represent the coupling degree of freedom, and the red arrows represent the
uniform tensile load, respectively

Fig. 11: Simulated displacement results of SRE sintered at (a) 275°C and (b) 300°C,
when subjected to a nodal tensile load of 1 MPa

P lx275 1MPa × 20µm


E275 = = = 17094.02MPa (4)
∆lx275 0.117 × 10−2 µm

P lx300 1MPa × 20µm


E300 = = = 29985.01MPa (5)
∆lx300 0.667 × 10−3 µm

∆ly275 0.377 × 10−3 µm


ν275 = − = = 0.3230 (6)
∆lx275 0.117 × 10−2 µm

16
Fig. 12: The probability density function curves of SRE with sintering temperature,
(a) equivalent Young’s modulus, (b) equivalent Poisson’s ratio

∆ly300 0.245 × 10−3 µm


ν300 = − = = 0.3682 (7)
∆lx300 0.667 × 10−3 µm

4.2 The Monte Carlo-based FE simulation and parameters


fitting
To scale the effect of the stochastically distributed voids on the dispersion of the
nanosilver sinter joint’s properties, as the method proposed in Section 2.3.2, 50 sets
of SREs are reconstructed based on the same pair of reconstruction parameters listed
in Table IV and imported into the Monte Carlo-based simulation process to get the
results of equivalent Young’s modulus and Poisson’s ratio.
The probability density function curves of the equivalent constitutive parameters
of SREs with sintering temperature are plotted in Fig. 12. Compared to the solid lines
representing the T s of 300°C, the dashed lines shift towards the left of the coordinate
axis, which means the equivalent Young’s modulus and equivalent Poisson’s ratio of
SREs decrease as the increased porosity of nanosilver sintered joint caused by the
lower sintering temperature. In addition, the distribution range of these constitutive
parameters at lower sintering temperatures is larger. It can be speculated that when
the sintering temperature goes down, the microscopic voids are distributed in a more
complex pattern due to the increased porosity and reduced average void size, as a
consequence, the dispersion of parameter variations improves.
With the highest acceptance probability values of 0.9944 and 0.9966, the prob-
ability density functions of the equivalent Young’s modulus are fitted into the
three-parameter Weibull distribution with the stochastic distribution form shown in
Eq. 8, and the fitted stochastic distribution for the equivalent Poisson’s ratio is the
Normal distribution with the function as Eq. 9. All the fitted parameters are listed in
Table VI.

17
 β−1 "  β #
β x−γ x−γ
f (x) = exp − (8)
η η η

where η is the scale parameter, β is the shape parameter and γ is the location
parameter.

"  2 #
1 x−µ
f (x) = √ exp − √ (9)
2πσ 2σ

where µ is the mean value and σ is the standard deviation.

Table VI The equivalent parameters fitting results


Sintering temperature (°C) 300 275 300 275
Parameter E ν
Distribution form Three-parameter Weibull Normal
ν/η 1.2133 2.6106 0.3632 0.3275
σ/β 2540 4150 0.008 0.0158
γ 26300 14000 / /
Acceptance probability 0.9944 0.9966 0.9670 0.9925

4.3 The multiscale modeling for nanosilver sintered joint


Referring to the design dimensions of the test samples, a macro scale simulation model
of the nanosilver sintered joint sample is built as shown in Fig. 13, where the block
from top to bottom is the silicon die, nanosilver sintered joint and copper substrate,
respectively. Considering the applied pressure and decomposition of organic matter
during the nanosilver sintering process, the modeled thickness of the sintered layer is
60 µm, and other dimensions are consistent with the values in Table I.

Fig. 13: The geometry model for macro finite element simulation

18
For this 3D model, the mesh element type is set as SOILID185 to simulate the
elastic behavior. Following the meshing requirements proposed in Section 2.5, the
element size of the sintered joint in X, Y, Z directions is determined as 20 µm, the
same as the SRE’s dimension. The same value is also used as the element size of the
silicon die and the copper substrate in X, Y directions to guarantee the continuous
nodes at the interfaces.
The material properties of the die and substrate and the density of the nanosilver
sintered joint are listed in Table I, while Young’s modulus and Poisson’s ratio of each
element in the sintered joint are sampled from the fitted distribution in last section.
Benefiting from this, the sintered joint is capable of characterizing the stochastic influ-
ence from micro voids without the need to model them physically. In this model,
187500 pairs of Young’s modulus and Poisson’s ratio values are sampled from the
Three-parameter Weibull distribution and the Normal distribution, respectively.

4.4 Nanosilver sintered joint FE simulation and experimental


verification
To verify the applicability of the FE model for nanosilver sintered joint built based
on the computational multiscale modeling method proposed in this paper, a FE sim-
ulation to get the shear strength as the experiment in Section 3.3 is developed as
follows.
Referring to the shear test conditions, the lower surface and the two sides along the
Y-axis direction of the substrate are set as completely fixed constraints to simulate the
fixture offered by a vacuum suction cup. A uniform load is applied on a side surface
of the die which is parallel to the X-Z plane with a coordinate of Y = -2.5 µm and the
direction is along the positive Y-axis., where the value of the load is calculated from
the critical shear force measured in the shear test as the Eq. 10.

Pshear = F/Across (10)

where P shear in MPa is the uniform load used in the simulation, F in N is the shear
force measured in the experiment and recorded in Table III, and Across represents the
area of surface that the shear load Pshear acting on, whose value is 300 µm2 .
In addition, the silicon die is set as a completely rigid body by coupling all degrees
of freedom of all its nodes, without considering the fracture of the silicon die during
shearing. The simulation boundary constraints and load conditions are shown in Fig.
14.
The simulation results of the stresses in the Y-Z plane of the nanosilver sintered
joint corresponding model to the sample 275-2 and the zoomed bonding layer are
shown in Fig. 15. It can be seen that the maximum shear stress in nanosilver sintered
joint occurs at the edge of the sintered layer with the greatest deformation during
the shear process. For other samples, the stress distribution patterns are analogous.
In this process, the shear deformation occurred in Y-Z plane, so with reference to
the factors of loading position, loading direction, and the shear strength calculation
method, the average shear stress response in Y-Z direction of the sintered layer nodes

19
Fig. 14: The (a) boundary conditions and (b) load of the macro finite element sim-
ulation model for nanosilver sintered joint, where the light blue arrows represent the
displacement constraints, the green arrows represent the coupling degree of freedom,
and the red arrows represent the uniform pressure load, respectively

Fig. 15: The simulation results of the shear stress in the Y-Z plane of (a) the sintered
sample 275-2 and (b) corresponding bonding layer

can be extracted as the simulated shear strength and compared with the experimen-
tal results from Section 3.3 and researches [48] and [49] where the samples are also
sintered under 1MPa, 275°C and 300°C, respectively, which is plotted in Fig. 16.
It can be seen that with the increase of the sintering temperature, the shear test
results and the finite element simulation results based on the computational multiscale
modeling method consistently showed the trend of increasing shear strength of the
joints; in addition, the simulation results are fairly consistent with the shear test
results, although there was a certain error. For the simulation, the error may have come
from the relatively small simulation input of Young’s modulus, which was obtained
from the nanoindentation test on the non-fully dense specimen.

20
Fig. 16: The simulated shear strength of the nanosilver sintered joints versus the
experimental results

Based on the classical double-sphere model [50] and Coble’s tetrahedron model
[51] describing the sintering process of material particles, elevating the sintering tem-
perature effectively promotes the sintering process and reduces the porosity of the
sintered body, creating more sintering necks and enriching the grain boundaries. The
presence of grain boundaries at room temperature can hinder the movement of dis-
locations, thus increasing the material’s resistance to plastic deformation. Therefore,
for nanosilver sintered joints, increasing the sintering temperature can enhance the
shear strength of the joints, consistent with the results of finite element simulations
conducted based on the multi-scale model proposed in this paper. Furthermore, for
nanosilver-sintered materials, strength and stiffness generally have a proportional rela-
tionship. Hence, nanosilver sintered joints obtained at higher sintering temperatures
have greater resistance to elastic deformation, and a higher Young’s modulus, which
shows consistent trends with the finite element simulation results of the nanosilver
sintered joint SRVE in microscale.
The above analysis declares that the simulation of nanosilver sintered joints based
on the computational multi-scale modeling method proposed in this study can prop-
erly predict the constitutive parameters distribution and the mechanical properties of
sintered joints, and the model can achieve more accurate prediction when the accuracy
of the input parameters is highly guaranteed.

5 Conclusions
A computational multiscale modeling method for nanosilver sintered joints with
stochastically distributed voids is originally proposed in this paper. Through the

21
finite element (FE) simulation on the micro-scale porous structure and the macro-
scale sintered samples with experimental verification, the following conclusions can be
drawn:
(1) A multiscale modeling method for nanosilver sintered joint is developed in 4
steps, including reconstruction of the stochastic representative elements (SREs) for
micro porous morphology, Monte Carlo-based FE simulation on SREs, distribution
fitting for constitutive parameters, and stochastically sampling for macro modeling.
This proposed method makes it possible to analyze the influence of the micro voids
on the macro properties in a simulation without modeling the micro details in the
simulation model.
(2) On the microscale, an approach to convert the morphology information of SRE
into an operational FE model is proposed where the SREs are reconstructed based on
the characteristics of the stochastically distributed voids. The tensile simulation on the
FE model shows that the stochastically distributed voids lead to the inhomogeneous
displacement distribution and the increased porosity of the sintered joints with lower
sintering temperature aggravates the deformation of corresponding SRE and worsens
the elastic tensile resistance.
(3) The distribution of equivalent constitutive parameters is captured from the
Monte Carlo-based FE simulation and parameters fitting and used as the link to
transfer the effect of the microscopic voids to the macroscopic scale of the sintered
joint. Influenced by the stochastic distribution of the voids, the equivalent Young’s
modulus shows a trend of three-parameter Weibull distribution and the equivalent
Poisson’s ration is Normal distribution. When the sintering temperature goes down,
the microscopic voids are distributed in a more complex pattern due to the increased
porosity and reduced average void size, which are observed in the experimental SEM
images and Monte Carlo-based simulation models, then the dispersion of parameter
variations improves.
(4) The proposed modeling method is used to simulate the shear process of the
nanosilver sintered samples and verified by the shear test, where the simulated and
experimental results of the shear strength show acceptable agreements, which turns
up the applicability of the proposed method. Both the simulation and the experiment
show that increasing the sintering temperature can improve the ability of nanosilver
sintered joints to resist shear damage, which is beneficial to reduce the possibility
of thermal fatigue damage in the solder layer of electronic packaging structures and
improve the inherent reliability level of nanosilver sintered materials.

Acknowledgments. The authors would like to thank Associate Professor Cheng


Qian and Associate Professor Hongqiang Zhang for their kind help in method
development and experiment, respectively.

Conflict of Interest. The authors declare that they have no conflict of interest.

Statements & Declarations


• Funding
This research is carried out as a part of the Center of Digitalized Electronics (CoDE)

22
project at Aalborg University funded by the Poul Due Jensen Grundfos Foundation
(Grant number 2021-001).
• Author Contributions
Zhongchao Sun: Conceptualization, Methodology, Software, Formal analysis and
investigation, Visualization, Writing - original draft preparation;
Wendi Guo: Conceptualization, Methodology, Software, Writing - review and edit-
ing;
Asger Bjørn Jørgensen: Visualization, Writing - review and editing, Funding
acquisition.
• Data availability
The data sets generated and analyzed during the current study are available from
the corresponding author upon reasonable request.
• Ethics approval
This research did not contain any studies involving animal or human participants,
nor did it take place on any private or protected areas.

Appendix A APDL code for building SRE FE model


The core code run in the ANSYS APDL environment for building SRE FE model is
shown below:
/ PREP7
! Define the model size
LENGTH =340
WIDTH =340
TOL =0.5 E -4
DOTNUMBER = LENGTH * WIDTH
MEDIUMNUMBER =25108
! Define the unit length
UL =0.0588
! Draw the key points
FLAG =1
i =1
* DOWHILE , FLAG
COLUMN = MOD (i , LENGTH )
* IF , COLUMN , EQ ,0 , THEN
COLUMN = LENGTH
* ENDIF
ROW =( i - COLUMN )/ LENGTH +1
K , ,( COLUMN -1)* UL ,( ROW -1)* UL ,0
K , , COLUMN * UL ,( ROW -1)* UL ,0
K , , COLUMN * UL , ROW * UL ,0
K , ,( COLUMN -1)* UL , ROW * UL ,0
i = i +1
* IF ,i , GT , DOTNUMBER , THEN
FLAG =0

23
* ENDIF
* ENDDO
! Load the . txt file recording the SRE morphology
* DIM , MEDIUM , ARRAY , MEDIUMNUMBER ,1 ,1
* VREAD , MEDIUM , MEDIUMFile , TXT
( F15 .0)
! Build the areas corresponding to the medium
FLAG =1
i =1
* DOWHILE , FLAG
p = MEDIUM ( i )
A ,4* p -3 ,4* p -2 ,4* p -1 ,4* p
i = i +1
* IF ,i , GT , MEDIUMNUMBER , THEN
FLAG =0
* ENDIF
* ENDDO
ALLSEL , ALL
ASEL ,S , AREA , ,1 , MEDIUMNUMBER ,1
CM , MEDIUMS , AREA
! Connect all separate medium ( sintered nanosilver )
ALLSEL , ALL
ASEL ,S , AREA , ,1 , MEDIUMNUMBER ,1 ,1
NUMMRG , KP

Appendix B APDL code for SRE FE simulation


The core code run in the ANSYS APDL environment for Monte Carlo-based FE
simulation of SREs is shown below:
/ PREP7
! Define the number of SREs
MODELNUMBER =50
! Calculate the equivalent Young ’ s modulus and Poisson ’ s ratio
* DO , model ,1 , MODELNUMBER ,1
! Build the SRE FE model
...
Code block 1
...
! Solve the model
/ SOL
ANTYPE , STATIC
! Define the fixed boundary
ALLSEL , ALL
NSEL ,S , LOC ,X ,0 - TOL ,0+ TOL
NPLOT

24
D , ALL , UX ,0
D , ALL , UY ,0
! Coupling the degree of freedom and Define the load
ALLSEL , ALL
NSEL ,S , LOC ,X , LENGTH * UL - TOL , LENGTH * UL + TOL
NPLOT
CP , NEXT , UX , ALL
CP , NEXT , UY , ALL
SF , ALL , PRES , -1
ALLSEL , ALL
SOLVE
FINISH
! Extract the results
! Calculate the equivalent Young ’ s modulus
! Calculate the equivalent Poisson ’ s ratio
...
Code block 2
...
! Clear current model and prapare for next model
/ PREP7
EDELE , ALL
NDELE , ALL
ADELE , ALL
KPDELE , ALL
* ENDDO

References
[1] Wang, Y., Ding, Y., Yin, Y.: Reliability of Wide Band Gap Power Electronic
Semiconductor and Packaging: A Review. Energies 15(18) (2022) https://doi.
org/10.3390/en15186670

[2] Jorgensen, A.B., Munk-Nielsen, S., Uhrenfeldt, C.: Overview of Digital Design
and Finite-Element Analysis in Modern Power Electronic Packaging. IEEE Trans-
actions on Power Electronics 35(10), 10892–10905 (2020) https://doi.org/10.
1109/TPEL.2020.2978584

[3] Lad, A.A., Hoque, M.J., Christian, S., Zhao, Y., Balda, J.C., King, W.P.,
Miljkovic, N.: High power density thermal management of discrete semiconduc-
tor packages enabled by additively manufactured hybrid polymer-metal coolers.
Applied Thermal Engineering 220(August 2022), 119726 (2023) https://doi.org/
10.1016/j.applthermaleng.2022.119726

[4] Suganuma, K.: Advances in lead-free electronics soldering. Current Opinion in


Solid State and Materials Science 5(1), 55–64 (2001) https://doi.org/10.1016/
S1359-0286(00)00036-X

25
[5] Abtew, M., Selvaduray, G.: Lead-free solders in microelectronics. Materials Sci-
ence and Engineering R: Reports 27(5), 95–141 (2000) https://doi.org/10.1016/
S0927-796X(00)00010-3

[6] Zhao, K., Zhao, J., Wei, X., Zhang, X., Deng, C., Yang, Y.: Mechanical
properties and microstructure of large-area diamond / silicon bonds formed
by pressure-assisted silver sintering for thermal management. Materials Today
Communications 34(December 2022), 105230 (2023) https://doi.org/10.1016/j.
mtcomm.2022.105230

[7] Yoon, J.W., Back, J.H.: Metallurgically and mechanically reliable microsilver-
sintered joints for automotive power module applications. Journal of Materials
Science: Materials in Electronics 33(3), 1724–1737 (2022) https://doi.org/10.
1007/s10854-022-07728-6

[8] Wang, C., Zhang, X., Zhang, Y., Zhao, T., Zhu, P., Sun, R., Nishikawa, H., Xu,
L.: Pressureless and low temperature sintering by Ag paste for the high temper-
ature die-attachment in power device packaging. In: 2022 IEEE 72nd Electronic
Components and Technology Conference (ECTC), vol. 2022-May, pp. 2256–2262.
IEEE, San Diego, USA (2022). https://doi.org/10.1109/ECTC51906.2022.00356

[9] Kobayashi, T., Ando, T.: Recent development of joining and conductive materials
for electronic components. Materials Transactions 62(8), 1270–1276 (2021) https:
//doi.org/10.2320/matertrans.MT-M2021060

[10] Chen, C., Kim, D., Zhang, Z., Wakasugi, N., Liu, Y., Hsieh, M.C., Zhao, S.,
Suetake, A., Suganuma, K.: Interface-Mechanical and Thermal Characteristics
of Ag Sinter Joining on Bare DBA Substrate during Aging, Thermal Shock and
1200 W/cm2Power Cycling Tests. IEEE Transactions on Power Electronics 37(6),
6647–6659 (2022) https://doi.org/10.1109/TPEL.2022.3142286

[11] Zhou, H., Guo, K., Ma, S., Wang, C., Fan, X., Jia, T., Zhang, Z., Xu, H., Xing, H.,
Wang, D., Liu, C.: A triple-layer structure flexible sensor based on nano-sintered
silver for power electronics with high temperature resistance and high thermal
conductivity. Chemical Engineering Journal 432(December 2021), 134431 (2022)
https://doi.org/10.1016/j.cej.2021.134431

[12] Chen, H., Zhang, J., Zhang, G., Liu, P.: Review of Laser Sintering of Nanosilver
Pastes for Die Attachment: Technologies and Trends. In: 2022 23rd International
Conference on Electronic Packaging Technology, ICEPT 2022. IEEE, ??? (2022).
https://doi.org/10.1109/ICEPT56209.2022.9872618

[13] Wang, J., Chen, S., Zhang, L., Zhao, X., Duan, F., Chen, H.: Brief Review
of Nanosilver Sintering: Manufacturing and Reliability. Journal of Electronic
Materials 50(10), 5483–5498 (2021) https://doi.org/10.1007/s11664-021-09078-1

[14] N’Tsouaglo, K.H., Milhet, X., Colin, J., Signor, L., Nait-Ali, A., Creus, J.,

26
Gueguen, M., Gadaud, P., Legros, M.: Time-Resolved Evolution of the 3D
Nanoporous Structure of Sintered Ag by X-Ray Nanotomography: Role of the
Interface with a Copper Substrate. Advanced Engineering Materials 24(1), 1–13
(2022) https://doi.org/10.1002/adem.202100583

[15] Wakamoto, K., Mochizuki, Y., Otsuka, T., Nakahara, K., Namazu, T.: Tempera-
ture dependence on tensile mechanical properties of sintered silver film. Materials
13(18) (2020) https://doi.org/10.3390/ma13184061

[16] Wereszczak, A.a., Vuono, D.J., Wang, H., Ferber, M.K.: Properties of Bulk
Sintered Silver As a Function of Porosity, pp. 1–36 (2012)

[17] Namazu, T.: Mechanical Property Measurement of Micro/Nanoscale Materials for


MEMS: A Review. IEEJ Transactions on Electrical and Electronic Engineering,
308–324 (2022) https://doi.org/10.1002/tee.23747

[18] He, G., Yao, Y., Yuting, Y.: Size effect on the fracture of sintered porous nano-
silver joints: Experiments and Weibull analysis. Journal of Alloys and Compounds
863, 158611 (2021) https://doi.org/10.1016/j.jallcom.2021.158611

[19] Wang, L., Ding, T., Gu, L., Sun, X.: Study on layer formation behavior of Ag
joints sintered with pressureless sintering process. Materials Research Express
9(11), 116512 (2022) https://doi.org/10.1088/2053-1591/ac9d84

[20] Wakamoto, K., Otsuka, T., Nakahara, K., Namazu, T.: Degradation mechanism
of pressure-assisted sintered silver by thermal shock test. Energies 14(17), 1–15
(2021) https://doi.org/10.3390/en14175532

[21] Chen, C., Zhang, Z., Kim, D., Sasamura, T., Oda, Y., Hsieh, M.C., Iwaki, A.,
Suetake, A., Suganuma, K.: Interface reaction and evolution of micron-sized Ag
particles paste joining on electroless Ni-/Pd-/Au-finished DBA and DBC sub-
strates during extreme thermal shock test. Journal of Alloys and Compounds
862, 158596 (2021) https://doi.org/10.1016/j.jallcom.2021.158596

[22] Yeo, S.M., Yow, H.K., Yeoh, K.H., Ishak, S.H.B.: Vacuum Reflow Process Opti-
mization for Solder Void Size Reduction in Semiconductor Packaging Assembly.
IEEE Transactions on Components, Packaging and Manufacturing Technology
12(8), 1410–1420 (2022) https://doi.org/10.1109/TCPMT.2022.3189995

[23] Gong, H., Wang, T., Zhu, J., Li, S., Yao, Y.: Compressive experimental analy-
sis and constitutive model of sintered nano-silver. Journal of Applied Mechanics
90(March), 1–29 (2022) https://doi.org/10.1115/1.4056253

[24] Xiong, Z., Wang, X., He, M., Benabou, L., Feng, Z.: Investigation on thermal
conductivity of silver-based porous materials by finite difference method. Mate-
rials Today Communications 33(June), 104897 (2022) https://doi.org/10.1016/j.
mtcomm.2022.104897

27
[25] Shen, Y.L., Abdo, M.G., Van Rooyen, I.J.: Numerical Study of Effective Thermal
Conductivity for Periodic Closed-Cell Porous Media. Transport in Porous Media
143(2), 245–269 (2022) https://doi.org/10.1007/s11242-022-01768-6

[26] Zhao, Z., Zhang, H., Zou, G., Ren, H., Zhuang, W., Liu, L., Zhou, Y.N.: A
Predictive Model for Thermal Conductivity of Nano-Ag Sintered Interconnect
for a SiC Die. Journal of Electronic Materials 48(5), 2811–2825 (2019) https:
//doi.org/10.1007/s11664-019-06984-3

[27] Renkuan, L., Hul, L., Haiyang, L., Wei, L., Xiao, W., Ran, Y., Renze, Y., Yue,
Y.: Analysis on the Reliability Effect of Solder Voids on the Nanosilver Sintered
Press-Pack IGBT. In: 2020 4th International Conference on HVDC, HVDC 2020,
pp. 20–25 (2020). https://doi.org/10.1109/HVDC50696.2020.9292846

[28] Qin, F., Hu, Y., Dai, Y., An, T., Chen, P., Gong, Y., Yu, H.: Crack Effect
on the Equivalent Thermal Conductivity of Porously Sintered Silver. Jour-
nal of Electronic Materials 49(10), 5994–6008 (2020) https://doi.org/10.1007/
s11664-020-08325-1

[29] Letz, S.A., Zhao, D., März, M.: Mesostructural impact on the macroscopic stress
state and yield locus of porous polycrystalline silver. Materials and Design 219,
110785 (2022) https://doi.org/10.1016/j.matdes.2022.110785

[30] Suzuki, T., Terasaki, T., Kawana, Y., Ishikawa, D., Nishimura, M., Nakako, H.,
Kurafuchi, K.: Effect of manufacturing process on micro-deformation behavior of
sintered-silver die-attach material. IEEE Transactions on Device and Materials
Reliability 16(4), 588–596 (2016) https://doi.org/10.1109/TDMR.2016.2614510

[31] Schaal, M., Klingler, M., Wunderle, B.: Silver Sintering in Power Electronics: The
State of the Art in Material Characterization and Reliability Testing. In: 2018 7th
Electronic System-Integration Technology Conference (ESTC), pp. 1–18. IEEE,
??? (2018). https://doi.org/10.1109/ESTC.2018.8546498

[32] Sun, Z., Wang, Z., Qian, C., Ren, Y., Feng, Q., Yang, D., Sun, B.: Characterization
of stochastically distributed voids in sintered nano-silver joints. In: 2019 20th
International Conference on Thermal, Mechanical and Multi-Physics Simulation
and Experiments in Microelectronics and Microsystems, EuroSimE 2019 (2019).
https://doi.org/10.1109/EuroSimE.2019.8724592

[33] Qian, C., Sun, Z., Fan, J., Ren, Y., Sun, B., Feng, Q., Yang, D., Wang, Z.: Char-
acterization and reconstruction for stochastically distributed void morphology in
nano-silver sintered joints. Materials and Design 196 (2020) https://doi.org/10.
1016/j.matdes.2020.109079

[34] Zou, Y.S., Gan, C.L., Chung, M.H., Takiar, H.: A review of interconnect materials
used in emerging memory device packaging: first- and second-level interconnect

28
materials. Journal of Materials Science: Materials in Electronics 32(23), 27133–
27147 (2021) https://doi.org/10.1007/s10854-021-07105-9

[35] Chen, C., Suganuma, K.: Microstructure and mechanical properties of sintered
Ag particles with flake and spherical shape from nano to micro size. Materials
and Design 162, 311–321 (2019) https://doi.org/10.1016/j.matdes.2018.11.062

[36] Li, S., Mao, Y., Liu, W., Hou, S.: A highly efficient multi-scale approach of
locally refined nonlinear analysis for large composite structures. Composite Struc-
tures 306(June 2022), 116578 (2023) https://doi.org/10.1016/j.compstruct.2022.
116578

[37] Alshahrani, A., Kulasegaram, S., Kundu, A.: Elastic modulus of self-compacting
fibre reinforced concrete : Experimental approach and multi-scale simulation.
Case Studies in Construction Materials 18(2022), 01723 (2023) https://doi.org/
10.1016/j.cscm.2022.e01723

[38] Hamed Mashhadzadeh, A., Fereidoon, A., Ghorbanzadeh Ahangari, M.: Com-
bining density functional theory-finite element multi-scale method to predict
mechanical properties of polypropylene/graphene nanocomposites: Experimental
study. Materials Chemistry and Physics 201, 214–223 (2017) https://doi.org/10.
1016/j.matchemphys.2017.08.042

[39] Costa, A.P.O., Seabra, M.R.R., Santos, A.D., Ribeiro, L.M.M., Cesar de
Sa, J.M.A.: Experimental and Numerical Multiscale characterization of a
Super Duplex Stainless Steel 25Cr-7Ni-Mo-N. Materials Today Communications
33(October), 104903 (2022) https://doi.org/10.1016/j.mtcomm.2022.104903

[40] Stefaniuk, D., Niewiadomski, P., Musiał, M., Łydżba, D.: Elastic properties of self-
compacting concrete modified with nanoparticles: Multiscale approach. Archives
of Civil and Mechanical Engineering 19(4), 1150–1162 (2019) https://doi.org/10.
1016/j.acme.2019.06.006

[41] Aduloju, S.C., Truster, T.J.: On topology-based cohesive interface element


insertion along periodic boundary surfaces. Engineering Fracture Mechanics
205(October 2018), 10–13 (2019) https://doi.org/10.1016/j.engfracmech.2018.10.
037

[42] Zhang, H., Wang, W., Bai, H., Zou, G., Liu, L., Peng, P., Guo, W.: Microstructural
and mechanical evolution of silver sintering die attach for SiC power devices
during high temperature applications. Journal of Alloys and Compounds 774,
487–494 (2019) https://doi.org/10.1016/j.jallcom.2018.10.067

[43] Tan, L., Liu, P., She, C., Xu, P., Yan, L., Quan, H.: Heat Dissipation Char-
acteristics of IGBT Module Based on Flow-Solid Coupling. Micromachines
(2022)

29
[44] Jorgensen, A.B., Cheng, T.H., Hopkins, D., Beczkowski, S., Uhrenfeldt, C., Munk-
Nielsen, S.: Thermal characteristics and simulation of an integrated GaN eHEMT
power module. In: 2019 21st European Conference on Power Electronics and
Applications. Institute of Electrical and Electronics Engineers Inc., ??? (2019).
https://doi.org/10.23919/EPE.2019.8915012

[45] Long, X., Tang, W., Xia, W., Wu, Y., Ren, L., Yao, Y.: Porosity and Young’s
modulus of pressure-less sintered silver nanoparticles. In: 2017 IEEE 19th Elec-
tronics Packaging Technology Conference, EPTC 2017, vol. 2018-Febru, pp. 1–8
(2018). https://doi.org/10.1109/EPTC.2017.8277577

[46] Yang, H., Wu, J.: Prediction of mechanical properties and fatigue life of nano
silver paste in chip interconnection. Research Square PREPRINT (2020)

[47] Zhang, H.Q., Bai, H.L., Jia, Q., Guo, W., Liu, L., Zou, G.S.: High Electrical
and Thermal Conductivity of Nano-Ag Paste for Power Electronic Applica-
tions. Acta Metallurgica Sinica 33(11), 1543–1555 (2020) https://doi.org/10.
1007/s40195-020-01083-3

[48] Lei, T.G., Calata, J.N., Lu, G.Q., Chen, X., Luo, S.: Low-temperature sinter-
ing of nanoscale silver paste for attaching large-area (>100 mm2) chips. IEEE
Transactions on Components and Packaging Technologies 33(1), 98–104 (2010)
https://doi.org/10.1109/TCAPT.2009.2021256

[49] Ide, E., Angata, S., Hirose, A., Kobayashi, K.F.: Metal-metal bonding process
using Ag metallo-organic nanoparticles. Acta Materialia 53(8), 2385–2393 (2005)
https://doi.org/10.1016/j.actamat.2005.01.047

[50] Kumar, V.: Simulations and Modeling of Unequal Sized Particles Sintering. PhD
thesis, The University of Utah (2011)

[51] Hsueh, H., Evans, A.G., Laboratorv, L.B., Engineering, M., Coble, R.L.: Develop-
ment during final intermediate stage sintering I. pore-grain boundary separation.
Acta Metallurgica 30, 1269–1279 (1982)

30
Figure 1 Click here to access/download;Figure;Figure 1.tif
Figure 2 Click here to access/download;Figure;Figure 2.tif
Figure 3 Click here to access/download;Figure;Figure 3.tif
Figure 4 Click here to access/download;Figure;Figure 4.tif
Figure 5 Click here to access/download;Figure;Figure 5.tif
Figure 6 Click here to access/download;Figure;Figure 6.tif
Figure 7 Click here to access/download;Figure;Figure 7.tif
Figure 8 Click here to access/download;Figure;Figure 8.tif
Figure 9 Click here to access/download;Figure;Figure 9.tif
Figure 10 Click here to access/download;Figure;Figure 10.tif
Figure 11 Click here to access/download;Figure;Figure 11.tif
Figure 12 Click here to access/download;Figure;Figure 12.tif
Figure 13 Click here to access/download;Figure;Figure 13.tif
Figure 14 Click here to access/download;Figure;Figure 14.tif
Figure 15 Click here to access/download;Figure;Figure 15.tif
Figure 16 Click here to access/download;Figure;Figure 16.tif
Graphical Abstract Click here to access/download;Graphical Abstract;Graphical Abstract.tif
Manuscript in PDF for reference

Click here to access/download


Supplementary Material
Revised_JEMS_Manuscript.pdf
Marked Manuscript in PDF for reference

Click here to access/download


Supplementary Material
Revised_Marked_Manuscript.pdf
Supplementary Material-Code

Click here to access/download


Supplementary Material
Supplementary files-APDL codes.docx

You might also like