Download as pdf or txt
Download as pdf or txt
You are on page 1of 77

Additive Manufacturing

Unraveling Jetting Mechanisms in High-Velocity Impact of Copper Particles Using


Molecular Dynamics Simulations
--Manuscript Draft--

Manuscript Number: ADDMA-D-23-01261R1

Article Type: Research Paper

Keywords: Jetting, High Velocity Impacts, Cold Spray, Shock Wave, Molecular Dynamics

Corresponding Author: Saeed Rahmati

CANADA

First Author: LM. Pereira

Order of Authors: LM. Pereira

A. Zúñiga

B. Jodoin

R. G. A. Veiga

Saeed Rahmati

Abstract: Jetting is an important phenomenon that takes place when a solid metallic particle
impacts a substrate at high velocity. It consists of material extrusion at the periphery of
the particle/substrate interface and is ubiquitous in particle deposition techniques such
as Cold Spray. Jetting is expected to play a key role in adhesion by removing the
native oxide layer and promoting metal/metal contact. In this study, large scale
Molecular Dynamics simulations are used to investigate at the atomic scale the
physical mechanisms of jetting formation during the impact of a copper particle with 0.2
μm diameter onto a copper substrate. Shock wave propagation and its effects on
deformation behavior are explained in detail. As the particle undergoes severe plastic
deformation, the localization of slip bands near the edges of the particle/substrate
interface significantly contributes to jetting onset. In addition, depending on the impact
velocity and resulting localized temperature increase, jetting can occur in either solid
phases (crystalline and amorphous) or molten phases, with the latter leading to the
ejection of material. Qualitative comparison with experimental data indicates that the
model, despite its size and approximations, captures a similar behavior experienced by
real particles.

Response to Reviewers: Dear Albert To,

Thank you for your email regarding the evaluation of our manuscript titled "Unraveling
Jetting Mechanisms in High-Velocity Impact of Copper Particles Using Molecular
Dynamics Simulations" (Manuscript Number: ADDMA-D-23-01261) submitted to
Additive Manufacturing.

We appreciate the reviewers' valuable feedback on our manuscript. We have carefully


considered their comments and have prepared a detailed response, addressing each
point raised. We have attached the response letter to the submission files, outlining the
changes made in response to their comments.

We are pleased to inform you that we have made the revisions requested by the
reviewers and have submitted the revised manuscript, and the response to reviewers’
comments.

We would like to express our gratitude to the reviewers for their insightful comments,
which have undoubtedly contributed to improving the quality of our work. We eagerly
await the reevaluation of our revised manuscript.

Thank you once again for your consideration and for valuing our contribution to
Additive Manufacturing.

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Kind regards,
Saeed Rahmati

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Response to Reviewers

Dear Reviewers,

I hope this finds you well.

I would like to begin by acknowledging the issues we encountered during the process of compacting
the figures in the PDF file for our manuscript. Unfortunately, these issues resulted in a reduction in
image quality. However, I want to assure you that the original figures were of good quality, and should
you choose to download them, you will see that the images are indeed fine. We sincerely apologize
for any inconvenience this may have caused. Despite our best efforts to address the problem by
adjusting the resolution, we were regrettably unable to resolve it on our end.

Please rest assured that we remain committed to upholding the highest standards of research and
presentation in our manuscript. We have taken into consideration all the valuable feedback provided
by the reviewers and have made the necessary revisions to enhance the clarity and scientific rigor of
the paper.

Thank you for your understanding and continued support. We appreciate the opportunity to address
these concerns and look forward to your favorable consideration of our revised manuscript.

Sincerely,
Saeed Rahmati
Reviewer 1: Unfortunately, we cannot address the comments of this reviewer as
she/he is clearly not referring to our manuscript.
Reviewer 2:

1. Page 6. Quasi-2D models with different diameter and the Brittle Oxide Layer are presented. But
only the 200 nm model is used in later writing, Brittle layer is also absent.

Authors' answers: Multi-million atoms MD simulations are costly in terms of required


computational resources and time-consuming. Therefore, the discussion on the effect of particle size
was concentrated in the "Particle Size Effects on Jetting" subsection of the manuscript, and
subsequent analyses focused on the largest particle considered in this study, which is the one closest
in size to real CS particles (this is indeed the reason we decided to use quasi-2D particles rather than
spherical ones, i.e., to be able to increase the particle size). It also exhibits more pronounced jetting
compared to smaller particles (please see Fig. 3(a-b)), with a morphology closer to what is observed
in FEM simulations and microscopy images. We would like to draw the reviewer's attention to the
discussion regarding the comparison of the simulated microstructure of the larger particle considered
in this study and a real CS particle, as illustrated in Figure 5.
To highlight even more the suitability of the 200 nm particle for further analysis, we added a sentence
where material ejection, which we observe in our simulations, was also observed in-situ:
“It is worth noting that Hassani-Gangaraj et al. also observed, in-situ, the ejection of jet-like material
after the supersonic impact of individual metallic microparticles [5]”

[5] M. Hassani-Gangaraj, D. Veysset, K.A. Nelson, C.A. Schuh, In-situ observations of


single micro-particle impact bonding, Scr. Mater. 145 (2018) 9–13.
https://doi.org/10.1016/J.SCRIPTAMAT.2017.09.042.

Regarding the brittle layer, it is always present; however, it is not typically shown in most of the
figures because our purpose is to highlight the jetting-related features by a number of different
analyses, such as CNA or DXA, where explicitly representing the brittle layer does not add useful
information. However, in Figure 3(c), the brittle layer (which is very thin) is indicated by arrows both
at the top and bottom of the deposited 200 nm particle. The observed gap in this simulation, which is
also observed in microscopy images of real powder particles used in CS, is a result of the brittle layer
not being removed in that particular region. To the best of our knowledge, this behavior is not
observed in MD simulations of particle impact where this brittle and chemically inert layer is not
present.
2. Page. 6 and 7. "MD simulations were carried out in order to evaluate the effect of particle geometry
on the deformation behavior of a spherical copper particle and a quasi-2D copper particle upon the
impact on a copper substrate at 1000 m/s.”
How is the spherical copper particle built in MD simulation? Please give proper reference or
description.
Why is 1000 m/s selected?
Authors' answers: We would like to thank the reviewer for point out that the description of the
spherical particle was missing in the manuscript. We added the following sentence for the spherical
particle:
“The spherical particle, also surrounded by a 0.5 nm thick brittle layer, was constructed following a
similar approach as described in Ref. [50].”

For the impact velocity, we clarified this by changing a sentence in the manuscript:
“During the rebounding process, the spherical particle remains bonded to the substrate, as the impact
velocity was above the critical velocity previously reported for a particle of same size [50].”

[50] L.M. Pereira, S. Rahmati, A. Zúñiga, B. Jodoin, R.G.A. Veiga, Atomistic study of
metallurgical bonding upon the high velocity impact of fcc core-shell particles, Comput. Mater.
Sci. 186 (2021) 110045. https://doi.org/https://doi.org/10.1016/j.commatsci.2020.110045.

3. Page 24. "In Fig. 8, it was shown that the shock wave reached to the particle top, and reflected back
towards the particle bottom as a tensile wave after 35 ps. This causes superpositions of larger tensile
waves, resulting in larger shear stresses and thus, increasing the shear deformation."
No shear deformation shown at the top. How the tensile wave reflected from top leads to shear
deformation?
To clearly see tensile wave, can tensile and compressive pressures be shown in different color contrast
in Fig. 8?
Authors' answers: The authors acknowledge that this sentence may lead to reader confusion and
express gratitude to the reviewer for bringing it to our attention. Upon further consideration, the
authors have decided to remove the sentence as it causes more confusion. The superposition of larger
tensile waves does not occur at 35 ps; instead, it takes place at a later time when the reflected wave
from the top reaches the bottom part of the particle.
The shock wave's intensity surpasses that of all other generated waves within the particle.
Consequently, the reflected wave from the top of the particle also exhibits high intensity. As this
reflected wave reaches the bottom part of the particle, it diminishes the hydrostatic pressure, enabling
easier deformation of the material. Additionally, through collision with other tensile waves in the
same region (i.e., the bottom of the particle), it creates a superposition of larger tensile waves that
facilitate shear strain in that area.
In Figure 8, an attempt has been made to illustrate the compression and tensile waves by plotting the
pressure distribution within the particle during deformation. It should be noted that there is a
substantial difference in intensity between the maximum tensile wave, the compression wave, and the
applied pressure. To provide some perspective, at a velocity of 700 m/s, the intensity of the applied
pressure is approximately 25 GPa, while the shock wave measures around 12 GPa, and the maximum
tensile wave reaches approximately 1 GPa. As a result, effectively depicting the wave propagations,
particularly after 35 ps when numerous collisions occur between different waves, becomes extremely
challenging.

4. Page 28. "Therefore, the shear strain pattern is independent from both microstructure and the
involved plasticity mechanisms, and is governed mainly by the applied shear stress which has an
arrow-head shape."
This paragraph should be the origin of the last point of Conclusion. But why just a reference is
provided? Any simulated or experimental proof? What are the 'plasticity mechanisms' here?

Authors' answers: We greatly value the reviewer's comment and would like to express our sincere
appreciation. It is essential to clarify that the state mentioned on page 28 was discussed after
establishing the consistency between our predicted results and the experimental observations
documented in the literature in the preceding sections.
To illustrate the similarity between our predicted results and experimental observations found in the
literature, we utilized a qualitative comparison depicted in Figure 5. Additionally, references
[7,53,58–60], on page 16 were referenced to further support the correspondence between our findings
and existing experimental data.
In references [31,32], we provided a comprehensive demonstration of the primary mechanisms
driving plasticity. These mechanisms encompass the nucleation and glide of Shockley partial
dislocations, dislocation multiplication and interlocking, as well as amorphization and dynamic
recrystallization during deformation.

[7] M. Walker, Microstructure and bonding mechanisms in cold spray coatings, Mater.
Sci. Technol. 34 (2018) 2057–2077. https://doi.org/10.1080/02670836.2018.1475444.
[31] S. Rahmati, A. Zúñiga, B. Jodoin, R.G.A. Veiga, Deformation of copper particles upon
impact: A molecular dynamics study of cold spray, Comput. Mater. Sci. 171 (2020).
https://doi.org/10.1016/j.commatsci.2019.109219.

[32] S. Rahmati, R.G.A. Veiga, B. Jodoin, A. Zúñiga, Crystal orientation and grain
boundary effects on plastic deformation of FCC particles under high velocity impacts,
Materialia. 15 (2021) 101004. https://doi.org/https://doi.org/10.1016/j.mtla.2021.101004.

[53] W. Xie, A. Alizadeh-Dehkharghani, Q. Chen, V.K. Champagne, X. Wang, A.T. Nardi,


S. Kooi, S. Müftü, J.-H. Lee, Dynamics and extreme plasticity of metallic microparticles in
supersonic collisions, Sci. Rep. 7 (2017) 5073. https://doi.org/10.1038/s41598-017-05104-7.

[58] T. Liu, J.D. Leazer, L.N. Brewer, Particle deformation and microstructure evolution
during cold spray of individual Al-Cu alloy powder particles, Acta Mater. 168 (2019) 13–23.
https://doi.org/https://doi.org/10.1016/j.actamat.2019.01.054.

[59] Y. Zou, W. Qin, E. Irissou, J.-G. Legoux, S. Yue, J.A. Szpunar, Dynamic
recrystallization in the particle/particle interfacial region of cold-sprayed nickel coating:
Electron backscatter diffraction characterization, Scr. Mater. 61 (2009) 899–902.
https://doi.org/10.1016/J.SCRIPTAMAT.2009.07.020.

[60] R. Nikbakht, M. Saadati, H. Assadi, K. Jahani, B. Jodoin, Dynamic microstructure


evolution in cold sprayed NiTi composite coatings, Surf. Coatings Technol. 421 (2021)
127456. https://doi.org/https://doi.org/10.1016/j.surfcoat.2021.127456.

5. Page 30. Please give color legend of Fig. 12 (b), (c), (d).
Authors' answers: The authors appreciate the reviewer for pointing out this mistake. The following
sentence was added to the Fig. 12’s caption:
(Different colors indicate different grains with low and high misorientation angle)
Reviewer 3:

1. The graphical abstract should be optimized. The figures and the labeling are difficult to read and
review.

Authors' answers: We would like to express our gratitude to the reviewer for bringing this mistake
to our attention. The graphical abstract has been recreated, and the font size of the labels has been
increased accordingly.
We would also like to acknowledge that there were some issues during the process of compacting the
figures in the PDF file, resulting in reduced image quality. However, we assure you that the original
figures had good quality, and if you download them, you will see that the images are fine. We
apologize for any inconvenience caused by this issue. Despite our efforts to resolve the problem by
adjusting the resolution, we were unable to resolve it on our end.

2. In section 2.0; the periodic boundary condition was only applied to the directions perpendicular to
the substrate. What are the boundary conditions for the other two axes?

Authors' answers: Indeed, as indicated in Fig. 1(a), periodic boundary conditions were applied along
X an Y axes. We acknowledge the reviewer for pointing out this mistake. We corrected the
corresponding sentence in the manuscript as follows:

“Periodic boundary conditions were applied to the directions parallel to the substrate surfaces, while
the direction perpendicular to the substrate surfaces was treated as non-periodic, as depicted in Fig.
1(a).”

3. What is the accuracy of the EAM and Lennard Jones potentials used in this study? Where has the
potential been used to model the deformation mechanism with the reference given? Please, provide
more information.

Authors' answers: About the potential for pure Cu by Mishin and co-workers, numerous properties
(lattice constants, thermodynamics, elastic, phonon curves) calculated for several crystal structures
can be found here: https://www.ctcms.nist.gov/potentials/entry/2001--Mishin-Y-Mehl-M-J-
Papaconstantopoulos-D-A-et-al--Cu-
1/EAM_Dynamo_MishinMehlPapaconstantopoulos_2001_Cu__MO_346334655118_005.html
Particularly for fcc Cu, the potential performs very well compared to ab initio calulations, which is
not surprising since data for this crystal structure (the most stable for Cu) were used in the fitting
procedure.

Regarding our work, the potential predicts that copper plasticity is governed by the glide of partial
Shockley dislocations, as expected. In this sense, the potential also performs quite well since the
intrinsic stacking fault energy, γSF, of copper predicted by the potential is close to the experimental
one, as summarized in Table V of Mishin’s paper (we are using potential EAM1 and also added this
information in the “Computational Approach” section to avoid ambiguity):

It is also worthwhile to mention that the paper presenting this potential has been cited more than 2,200
times on Google Scholar until May 2023.

Regarding the Lennard Jones potential, it serves as an approximation in the simulations to emulate
the role of the copper oxide layer in real powder particles. While this potential is too simple in its
formulation and does not capture the actual mechanical properties of copper oxide, it remains a
necessary approximation to replicate the observed behavior of particles during impact, as seen in
experimental observations. This is better justified in Ref. 50, where we first introduced this
approximation, but without incorporating this brittle layer, direct metal-to-metal contact would result
in adhesion regardless of the impact velocity, which is evidently inaccurate. The phenomena
demonstrated in Figures 2 and 3, i.e., particle rebounding below the critical velocity and the presence
of a gap at the center of the particle with adhesion occurring near the edges of the particle/substrate
interface (where jetting contributed to remove the brittle layer as it is expected to remove the oxide
layer in real CS particles), respectively, both of which have been experimentally observed, would not
be feasible without considering the presence of such a barrier hindering particle adhesion. Several
works that employ MD to simulate CS, including a recent publication in this journal [28], fail to
reproduce these fundamental aspects of the CS process by not considering this barrier to particle
adhesion.

[28] P. Gao, C. Zhang, R. Wang, G. Deng, J. Li, L. Su, Tamping effect during additive
manufacturing of copper coating by cold spray: A comprehensive molecular dynamics study,
Addit. Manuf. 66 (2023) 103448.
https://doi.org/https://doi.org/10.1016/j.addma.2023.103448.

4. In Fig. 1: The dimension of the substrate model should be given, particle speed should be indicated
and the fixed layers should be highlighted in another color.

Authors' answers: We would like to express our sincere appreciation for the reviewer's valuable
comment. It is indeed crucial to enable readers to obtain more valuable data from the figure and
enhance the overall comprehensibility of the modeling. We have taken all of your suggestions into
account and have made the necessary modifications to Figure 1 accordingly.

5. The paper studied shock wave formation and propagation inside a solid particle after impacting
onto a solid substrate at a high velocity. Is the particles' velocity a major factor? How were those
factors selected in this work?

Authors' answers: We appreciate this valuable comment. The impact velocity of particles plays a
significant role in generating shock waves, which in turn leads to plastic deformation in both the
particle and the substrate. Numerous studies have consistently demonstrated that the kinetic energy
of the particle upon impact is the primary factor influencing material deformation and deposition in
cold spray. It is well-established that kinetic energy is primarily derived from the mass and velocity
of the material.
In the section titled "Shock wave formation and propagation in particles," we investigated the effect
of different impact velocities ranging from 500 to 1000 m/s on jetting formation. At 500 m/s, no
jetting was observed in the particle. However, the first signs of jetting appeared at 600 m/s. The results
revealed that increasing the impact velocity also increased the occurrence of jetting.

6. A careful reading of the text should be done to suppress typo errors; can you check them, please?

Authors' answers: The authors appreciate this comment. The text has been carefully reviewed once
again, and the identified typos have been corrected.
Reviewer 4:

1. The graphic abstract is too blurry.. the resolution is too low..suggest changing

Authors' answers: We would like to express our gratitude to the reviewer for bringing this mistake
to our attention. The graphical abstract has been recreated, and the font size of the labels has been
increased accordingly.
We would also like to acknowledge that there were some issues during the process of compacting the
figures in the PDF file, resulting in reduced image quality. However, we assure you that the original
figures had good quality, and if you download them, you will see that the images are fine. We
apologize for any inconvenience caused by this issue. Despite our efforts to resolve the problem by
adjusting the resolution, we were unable to resolve it on our end.

2. With 83 references cited in this manuscript, there is no article cited from Additive
Manufacturing...please justify in the introduction how the study would benefit the additive
manufacturing community…

Authors' answers: We would like to thank the reviewer for reminding us of this mistake. We have
modified the Introduction of the paper and added new references that demonstrate the interest in Cold
Spray as an emerging additive manufacturing technology, the importance of understanding the
mechanisms underlying jetting formation for this technique, as well as the use of Molecular Dynamics
in the study of basic processes involved in additive manufacturing techniques, not only Cold Spray
but also others such as Selective Laser Melting. These references are recently published articles in
Additive Manufacturing.

3. The author has used MD to simulate particle impact (1000 m/s) on the substrate with a particle
diameter between 40-200 nm. However, the experiment results show a 12-um copper particle with an
impact speed of 650 m/s, it would be nice if they were compared within the same order..…

Authors' answers: We certainly agree that it would be really nice; however, as we mentioned in the
Introduction, the computational cost of considering particles of the same size as the powder particles
commonly used in Cold Spray in Molecular Dynamics simulations would be prohibitive. This
limitation is not specific to our work but also applies to other studies employing Molecular Dynamics
to investigate aspects of additive manufacturing, such as those now cited in the Introduction.
This is also the reason why we mentioned a qualitative comparison with experimental data in the
abstract. Achieving quantitative agreement, such as predicting the critical velocity of real copper
microparticles, is beyond the reach of atomistic simulations due to the difference in scale. In this work,
our ambition, and we believe we have been successful, is to reproduce microstructural features
observed in post-mortem microscopy analysis of particles deposited by Cold Spray, as shown in
Figure 5. The formation of metallic bonding near the periphery of the particle/substrate interface, with
a gap in the central region of the particle (Fig. 3(c)), is another experimentally observed feature that
our simulations were able to reproduce quite well. Finally, the formation of an amorphous layer at the
particle/substrate interface, observed in the MD simulations, was also observed experimentally in
Refs. [13,54–57].

[13] Y. Xiong, X. Xiong, S. Yoon, G. Bae, C. Lee, Dependence of Bonding Mechanisms of


Cold Sprayed Coatings on Strain-Rate-Induced Non-Equilibrium Phase Transformation, J.
Therm. Spray Technol. 20 (2011) 860–865. https://doi.org/10.1007/s11666-011-9634-0.

[54] S. Guetta, M.H. Berger, F. Borit, V. Guipont, M. Jeandin, M. Boustie, Y. Ichikawa, K.


Sakaguchi, K. Ogawa, Influence of particle velocity on adhesion of cold-sprayed splats, J.
Therm. Spray Technol. 18 (2009). https://doi.org/10.1007/s11666-009-9327-0.

[55] Y. Xiong, K. Kang, G. Bae, S. Yoon, C. Lee, Dynamic amorphization and


recrystallization of metals in kinetic spray process, Appl. Phys. Lett. 92 (2008) 194101.
https://doi.org/10.1063/1.2928218.

[56] Q. Wang, D. Qiu, Y. Xiong, N. Birbilis, M.-X. Zhang, High resolution microstructure
characterization of the interface between cold sprayed Al coating and Mg alloy substrate, Appl.
Surf. Sci. 289 (2014) 366–369. https://doi.org/https://doi.org/10.1016/j.apsusc.2013.10.168.

[57] K.H. Ko, J.O. Choi, H. Lee, Y.K. Seo, S.P. Jung, S.S. Yu, Cold spray induced
amorphization at the interface between Fe coatings and Al substrate, Mater. Lett. 149 (2015)
40–42. https://doi.org/https://doi.org/10.1016/j.matlet.2015.02.118.

4. Could you provide some experiment results of cold spraying Cu particles with particle size between
40-200 nm at 1000m/s-this is to help the reader to be convinced that the simulation results are
practical and reasonable in terms of process parameters.…
Authors' answers: We acknowledge the reviewer's suggestion regarding the experimental spraying
of Cu particles within the range of 40 to 200 nm for validating and verifying the MD results. However,
it is important to note that spraying such small particles with a CS (cold spray) system is impractical
due to the specific technological requirements of the CS apparatus. To the best of the authors'
knowledge, the spraying of copper particles of this size has not been carried out thus far.

5. In figure 2 (e-f), shows the dislocation network evolution. Could you provide some experimental
results to prove this is validated or cite some experiment results?

Authors' answers: The deformation process, and hence the evolution of the dislocation network,
occurs in a very short time—specifically, a few tens of nanoseconds. As a result, the in situ
observation of this process is nearly impossible with current technology.
Moreover, to our knowledge, there is no experimental study that has quantified the dislocation density
of an individual particle deposited by Cold Spray or obtained TEM images of the dislocation network
under this specific condition. This is certainly worth doing, but we understand that it may be
challenging experimentally. The results we know are only for complete coatings, as in Acta Materialia,
vol 13, 115-124 (2020). Additionally, the experimental samples may have undergone thermal or other
treatments, which would affect their microstructure. Therefore, direct comparison with these results
would not be meaningful.

6. Could you provide a table comparing the computer efficiency with running MD and current
published FEA simulation results.

Authors' answers: MD simulations typically require significantly more time and computational
resources compared to FEA. However, MD and FEA employ different approaches and have
different objectives. Therefore, comparing their computer efficiency is not meaningful as they serve
different purposes.
MD focuses on discrete mechanics, modeling atomic interactions, while FEA utilizes continuum
mechanics to model bulk materials. FEA is capable of modeling processes at actual size and time
scales, providing a macroscopic understanding of the system. On the other hand, MD is usually
unable to model real-sized processes too far from the nanoscale but offers valuable insights into the
underlying physical mechanisms at the atomic level.
While FEA lacks atomic detail, it can showcase behaviors based on the employed material model.
MD, in contrast, captures intricate atomic-level details, allowing for the identification of specific
physical mechanisms.
7. Some good experiment studies have been done to understand shock wave formation in other
area....(DOI 10.1088/0957-0233/23/12/125204).....it would be good if you could provide some
comparison to justify your simulation results and findings

Authors' answers: We appreciate the reviewer by suggesting to cite some other studies for better
justification of our results. We added the following sentence to the text:
“The strength of shock waves decreases as they propagate [68], as some of their energy is dissipated
as heat and is expended by displacing the atoms along their path.”
[68] P. Hough, T.J. Kelly, C. Fallon, C. McLoughlin, P. Hayden, E.T. Kennedy, J.P. Mosnier,
S.S. Harilal, J.T. Costello, Enhanced shock wave detection sensitivity for laser-produced
plasmas in low pressure ambient gases using interferometry, Meas. Sci. Technol. 23 (2012)
125204. https://doi.org/10.1088/0957-0233/23/12/125204.

8. There's no doubt that temperature plays a vital role in jetting; in your conclusion, could you
emphasize what the finding that separates between published FEA simulation results

Authors' answers: We consider that FEA and MD are both exceptional tools, but they do not rival
each other. They approach the phenomena of interest in different ways and, more importantly, they
usually do not answer the same questions. It is possible to simulate the same event, such as the impact
of a particle, and observe the same phenomena, such as jetting, using both FEA and MD. However,
MD provides a more fundamental physical understanding of observed phenomena direct from atomic
trajectories, while FEA, based on continuum physics, is limited in its ability to explain the underlying
mechanisms behind jetting formation precisely due to the absence of an atomic description.
It is true that FEA has greater predictive power precisely because FEA is performed on the same scale
of experiments and also takes the model parameters from experiments. On the other hand, MD
simulations, by integrating the equations of motion of discrete entities (atoms) that constitute the
system of interest, are capable of reproducing the behavior and features observed in real particles, as
we demonstrated in our manuscript, without the model requiring any experimental input. It only
requires a good description of atomic interactions in the potential formulation, which is adjusted based
on fundamental quantum mechanical calculations.
Highlights

1
2
3
4 Highlights
5
6
7
8
9
 Investigated the propagation of shock waves inside particles during high-velocity
10 impacts.
11  Explored the role of expansion waves (tensile waves) in plasticity.
12
13  Elucidated the mechanisms involved in the formation of jetting in Cold Spray.
14  Demonstrated the negligible role of "Hydrodynamic Pressure Release" mechanism in the
15 formation of jetting in Cold Spray.
16
17  Analyzed the mechanical and microstructural evolution during the jetting phenomenon.
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
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
Graphical Abstract Click here to access/download;Graphical Abstract;Graphical Abstract.tif

1
2
3
4
5
6
7
8
9
10
11
12
13
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
Fig. 1 Click here to access/download;Figure;Fig. 1.jpg
Fig. 2 Click here to access/download;Figure;Fig. 2.tif
Fig. 3 Click here to access/download;Figure;Fig. 3.tif
Fig. 4 Click here to access/download;Figure;Fig. 4.jpg
Fig. 5 Click here to access/download;Figure;Fig. 5.jpg
Fig. 6 Click here to access/download;Figure;Fig. 6.jpg
Fig. 7 Click here to access/download;Figure;Fig. 7.jpg
Fig. 8 Click here to access/download;Figure;Fig. 8.jpg
Fig. 9 Click here to access/download;Figure;Fig. 9.jpg
Fig. 10 Click here to access/download;Figure;Fig. 10.jpg
Fig. 11 Click here to access/download;Figure;Fig. 11.jpg
Fig. 12 Click here to access/download;Figure;Fig. 12.jpg
Author Statement

Author Statement

LM Pereira:
Investigation, Data Curation, Writing - Review & Editing

A Zúñiga:
Writing - Review & Editing

B Jodoin:
Funding acquisition, Supervision, Validation, Writing - Review & Editing

RGA Veiga:
Funding acquisition, Conceptualization, Supervision, Validation, Writing - Review & Editing

S Rahmati:
Conceptualization, Investigation, Data Curation, Supervision, Validation, Writing - Original
Draft, Writing - Review & Editing
Revised Manuscript (Clean Version) Click here to view linked References

Unraveling Jetting Mechanisms in High-Velocity Impact of Copper Particles Using Molecular

Dynamics Simulations

LM Pereiraa, A Zúñigaa, B Jodoinb, RGA Veigaa, S Rahmatic*

a
CECS, Universidade Federal do ABC, Santo André, SP, Brazil

b
Cold Spray Research Laboratory, University of Ottawa, Ottawa, ON, Canada

c
Centre for Advanced Coating Technologies, University of Toronto, Toronto, ON, Canada

* Corresponding author: saeed.rahmati@utoronto.ca

Abstract

Jetting is an important phenomenon that takes place when a solid metallic particle impacts a substrate

at high velocity. It consists of material extrusion at the periphery of the particle/substrate interface

and is ubiquitous in particle deposition techniques such as Cold Spray. Jetting is expected to play a

key role in adhesion by removing the native oxide layer and promoting metal/metal contact. In this

study, large scale Molecular Dynamics simulations are used to investigate at the atomic scale the

physical mechanisms of jetting formation during the impact of a copper particle with 0.2 μm diameter

onto a copper substrate. Shock wave propagation and its effects on deformation behavior are

explained in detail. As the particle undergoes severe plastic deformation, the localization of slip bands

near the edges of the particle/substrate interface significantly contributes to jetting onset. In addition,

depending on the impact velocity and resulting localized temperature increase, jetting can occur in

either solid phases (crystalline and amorphous) or molten phases, with the latter leading to the ejection

of material. Qualitative comparison with experimental data indicates that the model, despite its size

and approximations, captures a similar behavior experienced by real particles.

Introduction

Page 1 of 45
Additive manufacturing (AM) of complete 3D metal parts using powder particle deposition

techniques has garnered considerable attention from academia and industry in recent years. Cold

Spray (CS), in particular, stands out as highly suitable for this purpose. The CS process involves

accelerating solid micrometer-sized particles towards a substrate using a low-temperature supersonic

gas flow as the carrier. Upon impact, the particles experience severe plastic deformation at strain rates

of up to 109 s-1. CS has been widely applied to create metallic coatings across various industries, e.g.,

automotive, aerospace, and medical. Recent advancements in the technique have also allowed the use

of CS to build individual components with intricate geometries from the bottom up, layer by layer,

within a typical AM framework [1,2].

Numerous particle deposition studies at the microscale have demonstrated the occurrence of material

projection or ejection, commonly referred to as “jetting” [3,4], observed at the periphery of the

particle/substrate interface [5]. The jetting phenomenon is more observable as the particle impact

velocity increases. However, the nature of this phenomenon and its significance in particle deposition

is still a point of contention. Adiabatic shear instability (ASI) [6–10], and localized melting [11–13]

are the main mechanisms employed to explain the formation of jetting.

From a macroscopic point of view, the ASI mechanism occurs when the rate of thermal softening

exceeds the work hardening rate due to localization of plastic deformation. This local domination of

thermal softening causes a decrease in the flow stress, resulting in enhanced strain, and hence more

heat generation that further reduces the material strength. Under these conditions, material shows a

fluid-like flow behavior while remaining in the solid state. Conversely, localized melting can occur

at higher impact velocities if the generated heat is large enough to cause the material local temperature

to exceed its melting point, which can result in a significant deformation in the liquid state in that

region. Therefore, both ASI and localized melting mechanisms can be responsible for the formation

of jetting in CS.

In-situ observation of material behavior during CS impact is very challenging since the particle

deformation occurs over a very short time, i.e., tens of nanoseconds. Therefore, experimental

Page 2 of 45
observations usually take place after the spray is completed and rely on microscopy observations of

deformed particles’ cross-section. Attempts to explain the jetting formation based on the microscopic

images of the cross-section of deposited particles have been conducted [13–15].

Kim et al. [4] studied the jetting region in titanium splats using cross-sectional transmission electron

microscope (TEM) images. By finding very fine grains at the jetting zone, they suggested that due to

the occurrence of ASI in that region, the strain and temperature rose significantly, resulting in dynamic

recrystallization (DRX) of heavily deformed grains during the impact without any post-heat treatment.

Hassani-Gangaraj et al. [16] investigated the formation of jetting during microparticle impact in CS.

They asserted that because the shear bands, responsible for ASI, are rarely reported in the deposited

particles in CS, ASI is not responsible for material jetting. They proposed that the jetting occurs as

the result of strong shock waves interacting with the expanding edge of the particle and is formed as

a result of hydrodynamic plasticity. The hydrodynamic plasticity term is used when the pressure

generated by a shock wave exceed the material strength by at least one order of magnitude, resulting

in the solid material beginning to response according to hydrodynamics laws [17–19], i.e., an

instability state. Rayleigh-Taylor and Richtmyer-Meshkov instabilities are two main theories that

have been used as effective tools to predict the material behavior under shock waves that are above

the Hugoniot Elastic Limit (HEL) [20–23]. However, these continuum-based theories are unable to

explain the microstructural evolution that happens during the formation of jetting.

Zasimchuk et al. [24–27] studied material deformation behavior under highly dynamic conditions.

They showed that any sudden changes in the external mechanical field, which creates a

thermodynamic non-equilibrium condition, soften the material significantly and causes an increase

in plasticity. They showed that the material softening is due to the creation of micro bands and/or

shear bands, regardless of the increase in the external stress. They demonstrated that the localization

of bands increases the vacancy defects localization, resulting in a liquid-like structure inside the solid

structure (hydrodynamic channel). Therefore, the plastic deformation will be concentrated inside the

Page 3 of 45
hydrodynamic channels, as these channels have a very low shear resistance compared to other regions.

This behavior was called hydrodynamic flow channel and it can explain the ASI mechanism.

An alternative to investigate the CS process from a more fundamental perspective, in an attempt to

address questions that are difficult to tackle even with the most advanced in-situ experiments, is the

computational atomic scale technique known as Molecular Dynamics (MD). Gao and co-workers, for

instance, applied MD simulations to study the so-called tamping effect during CS additive

manufacturing of copper coatings [28]. Also, in the field of AM, MD simulations of selective laser

melting [29] and sintering [30] provided valuable insights into the fundamental mechanisms

underlying these manufacturing processes.

In previous studies, the deformation behavior of face-centered cubic (FCC) nanoparticles during high

velocity impact were investigated using MD. It was shown that a particle with 40 nm diameter is large

enough to capture the mechanical responses of material similar to what is observed during the

deformation of micron-size particles [31]. The lattice orientation and grain boundary effects on plastic

deformation were studied, and it was demonstrated that plastic deformation in FCC nanoparticles at

high velocity impacts involve the formation of Low-angle Grain Boundaries (LAGBs), High-angle

Grain Boundaries (HAGBs), twining, and grain rotation [32]. Furthermore, it was demonstrated that

a significant proportion of atoms (over 30 percent) experience a loss of their crystalline structure

during the impact. Consequently, solid-state amorphization was identified as one of the two primary

mechanisms contributing to the plastic deformation observed in these particles. These findings were

consistent with the hydrodynamic flow channel mechanism suggested by Zasimchuk et al. [24].

However, it is important to note that jetting in these nanoparticles was not as evident as jetting in CS

microparticles observed by microscopy [3,15] and reproduced in finite element method (FEM)

simulations [33,34]. This suggests a significant influence of particle size on the MD results.

Despite being undoubtedly a powerful tool to investigate the physical grounds of plasticity at the

atomic scale, MD has important limitations as to the size of the systems that can be studied. Typically,

MD simulations with up to a few tens or even hundreds of millions of atoms are feasible in today's

Page 4 of 45
high-performance computing systems. However, a single micrometer-sized particle, comparable to

the powder particles used in CS, can be made up of hundreds of billions of atoms and is therefore far

beyond the reach of MD.

To mitigate MD size limitations and investigate the deformation behavior (with a focus on jetting

formation) of particles with a size closer to the microscale, a quasi-two-dimensional (quasi-2D)

system is considered in this study. Dimension reduction, i.e., from three to two dimensions, allows

MD simulations of systems much larger in the remaining two dimensions while the system retains

some meaningful physical properties. However, suppressing a dimension is usually not possible

except for MD simulations that use simple pairwise potentials, such as Lennard-Jones potential. This

dimension reduction is then accomplished in a system that remains three-dimensional, where the

dimension to be reduced is very thin compared to the others, and periodic. As such, the system, which

is usually finite or contains interfaces, is forced to present a two-dimensional symmetry in the plane

perpendicular to the reduced dimension. For example, quasi-2D polycrystalline systems with

idealized columnar microstructures were used to study dislocation processes [35] and grain growth

[36]. More recently, Kurian and Mirzaeifar also used a quasi-2D approach with MD to study the

selective laser melting of aluminum powder particles [37]. The quasi-2D particles in this study are

constructed as cylinders with infinite length along one dimension. In the current case, it enables to

model particles with diameters up to 0.2 µm. This particle diameter is expected to be large enough to

capture the deformation behavior experienced by particles used in CS while keeping the number of

atoms within the practical limits of MD. The deformation behavior of a cylinder impacting a substrate

perpendicular to its axis is expected to be somehow affected by this particular geometry when

compared to the impact of a spherical particle. This point is addressed by direct comparison of a

quasi-2D particle with a spherical particle. Moreover, there are important similarities that demonstrate

why 2D systems have been widely used to study the particle impact and the jetting formation [38–

46]. In addition, Field et al. [47] verified that although the pressure field is higher in a cylindrical (or

Page 5 of 45
2D) drop than a spherical drop, the shock waves’ propagation and interaction mechanisms are

identical for both geometries.

The main objective of this study is to use MD to reveal the physical nature of jetting and material

ejecta which are common features observed in high velocity impacts of microparticles. In this work,

the shock wave propagation during the high velocity impacts and its role on the formation of jetting

is uncovered. In addition, the basic plasticity mechanisms, i.e., nucleation, multiplication, gliding,

interaction of dislocations, and amorphization during the formation of jetting are investigated.

Computational approach

Molecular Dynamics simulations were performed using the LAMMPS code [48] with a time step of

1 fs for the integration of Newton equations of motion of a system of point particles representing the

atoms. Periodic boundary conditions were applied to the directions parallel to the substrate surfaces,

while the direction perpendicular to the substrate surfaces was treated as non-periodic, as depicted in

Fig. 1(a). The equilibration at 300 K was carried out in the NVT (i.e., number of particles, volume

and temperature held constant) ensemble, switched to the NVE (i.e., number of particles, volume and

energy held constant) ensemble in the impact simulations, where particles were given a prescribed

translational velocity normal to the substrate. The systems used in the MD simulations consisted of

particles with diameters of 40, 80, 120 and 200 nm initially positioned 1 nm above the substrates. The

thickness of the reduced, periodic dimension, aligned with the cylinder axes, was 7 nm. In each system,

both the particle and the substrate were made up of copper atoms arranged on a face centered cubic

(FCC) lattice with a lattice parameter equal to 0.361 nm. To prevent substrate displacement after

impact, three atomic layers at the substrate bottom were fixed in impact simulations. Interactions

between metallic atoms were described by the embedded atom method (EAM) potential for elemental

copper developed by Mishin and co-workers [49] (labeled EAM1 in the reference). In order to mimic

to some extent the role of the oxide layer which is ubiquitous in almost all metallic powder particles,

a 0.5 nm thick outer layer in the particles has been defined as a brittle and chemically inert layer

according to the model introduced in Ref. [50]. Lennard-Jones potentials were employed to describe

Page 6 of 45
the interactions between metallic atoms inside the particle and atoms in the brittle layer, as well as

for the interactions of the brittle layer atoms with each other and the interactions of those atoms with

the metallic atoms in the substrate. Ref. [50] contains the values of the Lennard-Jones parameters for

each type of interaction. Fig. 1(b) schematically depicts the 40, 80, 120, and 200 nm particle-substrate

system, and the total number of atoms of each system.

Figure 1 a) Model system and the direction of the periodic boundary conditions, b) Schematic

representation of 40, 80, 120, and 200 nm particle-substrate systems, and the total number of atoms

in each system

Results and Discussion

Validation of Using Quasi-2D Particles in Studying the Jetting Phenomenon

MD simulations were carried out in order to evaluate the effect of particle geometry on the

deformation behavior of a spherical copper particle and a quasi-2D copper particle upon the impact

on a copper substrate at 1000 m/s. The spherical particle, also surrounded by a 0.5 nm thick brittle

layer, was constructed following a similar approach as described in Ref. [50]. The diameter of both

particles is 40 nm, which is a practical limit for the spherical particle in atomistic simulations [31].

Fig. 2(a) and (b) display the temperature distribution on the cross section of the spherical and the

quasi-2D models, respectively, as well as the changes in particle shape after impacting the substrate.

In order to illustrate the similar behavior between the quasi-2D and the spherical models, different

Page 7 of 45
scales were selected for this figure. Despite similarities in temperature distribution, the maximum

temperature in the quasi-2D particle is less than the spherical particle, and also the flattening of the

latter is more pronounced. The generated heat for each particle follows the shear strain pattern, which

was shown in [32]. For both particles, at the early stage of the impact process, the deformation is

confined at the bottom and subsequently develops towards the center of the particles. At around 20

ps, the onset of the material jetting formation can be observed. At the late stage of the deformation,

the particle’s kinetic energy decreases significantly, resulting in a drop of strain rate. By decreasing

the local deformation rate, the heat evacuation rate across the particle became noticeable. The

temperature at the contact area and the jetting zone decreases and heat flows to the center and the top

part of the particle, a seen at 25 and 50 ps in Fig. 2.

Page 8 of 45
Figure 2 (a-b): Temperature distribution, (c-d): Atomic structure (green atoms are FCC, red atoms

are HCP and white atoms are non-crystalline), and (e-f): Dislocation network evolution during the

impact for the spherical and the quasi-2D models. (a, c, and e) belong to the spherical model, and

(b, d, and f) belong to the quasi-2D model.

Fig. 2(c-f) shows the atomic structure and dislocation network evolution of the spherical and the

quasi-2D particles during the impact. The same mechanisms are involved in the deformation of both

particle models upon high velocity impact, i.e., dislocation emission and glide, with formation of a

dense dislocation network, beginning in the contact zone towards the center of the particle, and solid-

state amorphization concentrated in the jetting zone. However, due to the differences between the

Page 9 of 45
models’ geometry, i.e., a sphere and an infinite cylinder, the mechanisms’ contributions and behaviors

are different, which leads to some alterations to the particles’ final shape and captured mechanical

responses during the impact. For instance, the maximum fraction of non-crystalline atoms (or

amorphous structure) during the impact process reaches to 55.1% and 32.6% in the spherical and

quasi-2D models, respectively. In a previous study [32], it was shown that the solid-state

amorphization during the impact process is induced by a high localized dislocation density. The

dislocation multiplication rate for the spherical model is higher compared to the quasi-2D model. In

addition, the larger amount of amorphization for the spherical model might be the reason for the

higher temperature in this model, especially close to the center of the particle, see Fig. 2.

Due to the large deformation at the contact area, the brittle layer failed for both particles and allowed

the formation of temporary metallic bonding between atoms of the particle and the substrate. During

the rebounding process, the spherical particle remains bonded to the substrate, as the impact velocity

was above the critical velocity previously reported for a particle of same size [50]. On the other hand,

the quasi-2D model was detached completely from the substrate at the same impact velocity. This

indicates that the predicted critical velocity is higher for the quasi-2D models.

In conclusion, the comparison between the quasi-2D model and the spherical model indicates that the

involved plasticity mechanisms are similar in both models. As demonstrated in this section, although

the amount of plastic deformation in the quasi-2D model was less than the spherical model, the

deformation pattern was similar for both models. The quasi-2D model presented the jetting

phenomenon which is the main subject of this study. Using the quasi-2D system allows to model

larger particles, and discovering the formation of jetting in a scale closer to that of real particles.

Particle Size Effects on Jetting

By reducing the dimensionality of the impacting particle model, it is possible to simulate particles

with much larger diameters than spherical particles with MD. In order to evaluate the influence of the

Page 10 of 45
particle size on the deformation behavior, quasi-2D copper particles with diameters of 40, 80, 120,

and 200 nm impacting copper substrates at 1000 m/s were studied.

Temperature distribution of particles at the end of deformation is shown in Fig. 3(a). At the same

impact velocity, the larger particles (having more kinetic energy upon impact to be redistributed)

present more plastic deformation in the zone near the substrate, which in turn resulted in more heat

generation in the localized zone resulting in a rise of the material temperature especially near the

contact zone and the jetting area. The limited material jetting for both 40 and 80 nm particle occurred

without the material reaching the melting point: all deformation occurred in the solid-state. The

maximum calculated temperatures were 750 K and 1050 K for 40 and 80 nm particles, respectively.

However, for larger particles, i.e., 120 and 200 nm, the material temperature exceeded the melting

point at the tip of the jetting area, causing material ejecta in the molten-state (or liquid-state) in these

particles. It is worth noting that Hassani-Gangaraj et al. also observed, in-situ, the ejection of jet-like

material after the supersonic impact of individual metallic microparticles [5].

Page 11 of 45
Figure 3 a) Temperature distribution after the impact at 1000 m/s, b) Final deformed shape and

atomic structure (green atoms are FCC, red atoms are HCP and white atoms are non-crystalline), c)

Enlargement of the particle/substrate interface periphery, illustrating the bonding area, the

remaining brittle layer, and the gap between the particle and the substrate, and d) Fraction of non-

crystalline atoms during plastic deformation for 40, 80, 120, and 200 nm particle.

The final deformed shape of each particle is shown in Fig. 3(b). After the impact, the brittle shell

surrounding the particles broke and was partially removed near the edges of the particle-substrate

interface, see Fig. 3(c). As a consequence, in that region, the particles metallic core was put in direct

contact with the metallic substrate, resulting in the formation of metallic bonds. However, although

metallic bonding is a well-known mechanism for the adhesion of metallic particles in CS, it has been

Page 12 of 45
shown both experimentally [51–53] and from simulations [33,34,50] that it may not be enough to

prevent particle detachment when restoration of elastic energy occurs (rebounding). From the

simulations, the impact velocity was lower than the critical velocity for the case of quasi-2D 40 nm

particle. This particle bounced-off from the substrate after breaking the metallic bonds that were

established with the substrate upon impact. In contrast, for the 80, 120 and 200 nm particles, the

impact velocity was high enough to allow the particles to adhere sufficiently to the substrate, i.e., the

number of metallic bonds formed upon impact was sufficient to keep them attached when the particles

started moving upwards.

As reported previously in MD simulations of nanoparticles impacting a substrate [31,32,50], the

formation of solid-state portions of amorphous material is one of the two main mechanisms (the other

is dislocation creation and glide) involved in the severe deformation of metallic particles upon high

velocity impact. The destruction of the crystalline structure is mostly concentrated near the particle-

substrate contact region, particularly in the jetting zone. Fig. 3(d) displays the evolution of the fraction

of non-crystalline atoms throughout the MD simulations. Interestingly, there is a clear size-dependent

trend: the maximum degree of amorphization is consistently reduced as the particle diameter increases.

The decrease in the fraction of non-crystalline atoms can be due to the fact that the amorphization

mainly happens at the jetting zone, and obviously, the ratio between the number of atoms that are in

the jetting zone over the total number of atoms in the particle decreases by increasing the particle size.

For the largest particle in this study, with a diameter of 0.2 μm particle, the maximum fraction of non-

crystalline atoms was 28%. If this trend were maintained, it should be expected that the degree of

amorphization of microparticles, which is the typical size range of CS-deposited particles, impacting

a substrate also at 1000 m/s would be significantly lower than the 0.2 μm particle. This might be the

reason that the solid-state amorphization mechanism is hardly observable from microscope images

taken of the jetting zone in microparticles. However, this mechanism can play an important role for

the formation of jetting through the creation of hydrodynamic channels [24–27]. Also noteworthy, for

the 40 nm particle, the fraction of non-crystalline atoms reached a maximum, about 32.6%, and then

Page 13 of 45
decreased, eventually going below the value of the 0.2 μm particle after around 120 ps. This drop in

the fraction of non-crystalline atoms was related to the subsequent crystallization of amorphous

material when heat started to be distributed inside the particle and in the substrate, a feature already

observed for a 40 nm spherical particle in [50]. The same phenomenon of formation of non-crystalline

material also took place for the biggest particles, for which the fraction of non-crystalline atoms rather

reached a plateau and remained there for the rest of the simulation. In these particles, unlike the 40

nm particle, the temperature in some regions, particularly in the jetting zone, rose above the melting

point of copper, so part of the non-crystalline atoms corresponded, in fact, to molten material.

The decrease in the percentage of amorphization by increasing the particle size can be observed

visually in Fig. 4. In this figure, the atomic structure and the dislocation network at the time that the

fraction of non-crystalline atoms was maximum for each particle are shown. The white atoms in the

in Fig. 4(a-b), and the gray area and also zones with no dislocations in Fig. 4(c) illustrate the generated

amorphous structure in the particles during the deformation. For the 40 nm particle, the amorphous

structure occupied a thick portion of the atomic structure, especially in the jetting zone. By increasing

the particle size, the allotment of amorphous structure reduced and the amorphous region became

thinner. For the 0.2 μm particle, localized amorphous structure was only found at the jetting tip and a

very thin layer at the particle bottom. The generation of a thin layer of amorphous structure at the

interface between the particle and the substrate has been observed experimentally in [13,54–57]. The

rest of amorphous structure were mostly either LAGBs or HAGBs, which were formed during the

plastic deformation, as already discussed in [32].

Page 14 of 45
Figure 4 a) Atomic structure (green atoms are FCC, red atoms are HCP and white atoms are non-

crystalline), b) enlargement of selected areas in (a), and c) dislocation network of (b)

Figs. 5(a) and (b) display respectively an ion beam image and an electron beam image of the cross

section of a 12 μm copper particle deposited to a copper substrate at around 650 m/s [15]. Figs. 5(c)

and (d) show different magnification of the MD result of a 0.2 μm copper particle at 300 ps after the

impact to a copper substrate at 1000 m/s. In [15], it was shown that the average deformation in the

selected regions in Fig. 5(a) increased by moving from particle top to particle bottom (or from A to

D), and maximum deformation was found in region E, which is the area close to jetting zone. The

MD-predicted microstructure after impact of the quasi-2D particle reproduces the main

microstructural features observed in the experimental data despite the differences in size and

geometry. The formation of HAGBs at the interface indicates the larger deformation which occurred

Page 15 of 45
in this area compared to the top of the particle. In addition, similar to the experimental results shown

in Fig. 5(a-b) and also other experimental studies [7,53,58–60], a large number of smaller grains can

be observed near the jetting zone, which demonstrate the high amount of dynamic recrystallization in

that area. This qualitative comparison between the microscopy observation and the numerical results

provides additional evidence that the MD simulation of the impact of the 0.2 µm quasi-2D particle

captures a similar behavior compared to actual microparticles in CS. Thus, it gives some more

confidence in using a quasi-2D model to perform atomic scale investigation of high velocity single

particle deposition.

Figure 5 a) Ion beam image, and b) Electron beam image of a 12 µm copper particle impacted to a

copper substrate at around 650 m/s. c) Final deformed shape of a 0.2 μm copper particle at 300 ps

after impacting to a copper substrate at 1000 m/s (black color shows interface and HAGBs, gray

show FCC structure, and white shows defects like bands and/or stacking faults), d) Higher

magnification of the particle/substrate right edge. (a-b) Images after King et al. [15], and (c-d)

Images representing numerical results obtained through CNA analysis.

Page 16 of 45
Shock wave formation and propagation in particles

Apart from the high velocity deposition of solid particles, jetting has also been reported and discussed

in other processes such as explosive welding [61], droplet impact [62], and shaped charges [63]. The

generation and propagation of shock waves is a common feature among all these processes and is

credited as the main reason behind severe material deformation. In CS, a large number of infinitesimal

compression waves are generated at the impact point due to the abrupt deceleration of the particle

upon collision with the substrate. These infinitesimal compression waves quickly coalescence into a

finite-strength shock wave. Wave propagation in a material depends on both its geometry and its

mechanical properties. In order to understand this phenomenon, the creation, propagation, and

interaction of waves inside the solid particles will be first examined in this section.

A schematic, bidimensional representation of wave propagation inside a particle is shown in Fig. 6(a)

and henceforth it will be used to support the results obtained and the deformation behavior of the

simulated particles. The quasi-2D particle is depicted as non-deformed in Fig. 6(a) for simplicity.

While it is well-known that an infinite number of compression waves are generated during the impact,

only the propagation of a few waves and their resulting coalescence into a finite strength shock wave

are shown, Furthermore, waves interact with each other when they collide but these interactions

between the waves are not shown in this figure also for simplicity.

As soon as the particle impacted the substrate, each atom hitting the substrate generate an

infinitesimal compressive wave which travels towards the particle top at the local speed of sound, or

a continuous train of compressive waves as illustrated in Fig. 6(a) at 𝑡 = 𝑡1 . Each infinitesimal

compressive wave increases the material’s density and temperature along its path, causing an increase

in the sound velocity of that condensed matter. Thus, each infinitesimal compressive wave travels

faster than the prior wave generated previously, and hence they coalesce together shortly after the

impact, and move together as a “shock wave” (the blue curve in Fig. 6(a) at 𝑡 = 𝑡2 ). When the shock

wave reaches the free surfaces (i.e., surfaces which are not in contact with the solid substrate), tensile

(or expansion) waves, the red curves in Fig. 6(a) at 𝑡 = 𝑡3, reflect back towards the particle center

Page 17 of 45
while the shock wave continues towards the particle top. Tensile waves are generated in order to

satisfy the stress-free boundary conditions at free surfaces. At time 𝑡 = 𝑡4 , the tensile waves meet

each other, resulting a larger stress at their collision zone. At time 𝑡 = 𝑡5 , the two tensile waves

separate and they continue travelling forward [64]. After 𝑡5 , the blue compression wave reaches the

particle top free surface and hence, it reflects back towards the particle bottom in the form of tensile

wave. Accordingly, the red tensile waves will also reflect back in the form of compression waves,

when they reach the free surfaces.

Page 18 of 45
Figure 6 a) Schematic of shock wave formation and propagation inside a solid particle after

impacting onto a solid substrate at a high velocity. Blue and red colored lines display the shock

wave and reflected (expansion or tensile) waves, respectively. b) Schematic representation of

pressure distribution in a particle after impacting to a substrate at high velocities

All generated waves interact simultaneously with each other, and raise or destroy each other

depending on their types. Besides the interaction between waves, each wave will reflect back in its

opposite type by reaching the particle free surfaces and even the particle-substrate interface. The

amount of the wave reflection at the particle-substrate is dependent on the gap between the particle

and the substrate, and also the density of the matter at the incident time [65]. By considering all

mentioned phenomena, the particle can be seen as a zone of various waves that propagate in different

directions, interact with each other concurrently and promote or terminate each other continuously.

Fig. 6(b) displays a schematic representation of pressure distribution inside a particle after impacting

on a substrate. Three areas are shown in this figure. First, a shock wave, resulting from the rapid

coalescence of the compression waves generated upon impact of the particle, which is rapidly

propagating towards the top part of the particle. Second, tensile waves propagation areas which are

zones that were affected by the propagation of the reflected compression waves into tensile waves.

The propagation of tensile waves reduces the shock wave effect by lowering the pressure in that

region, see Fig. 6(b), and third, the hydrostatic pressure region that occurs at the impact point. Upon

the impact, a high pressure is generated at the particle south pole, as shown in Fig. 6(b). The material

in this area has not seen the expansion waves, and is restricted to move in any directions, thus the

name “hydrostatic pressure”.

Metals with crystalline structure are composed of atoms in ordered layers [66,67]: pressure in a solid

metal is generated when the distance between these atomic layers is reduced. After the passage of the

shockwave, the distance between the various atomic layers keeps decreasing as they slow down and

eventually reach zero velocity, In Fig. 6(b), it can be seen that there is a gradient of pressure behind

Page 19 of 45
the shock wave. This pressure gradient is formed by the velocity gradient that exist between the

atomic layers after the shockwave passage.

Aiming to explore the nature of jetting and the mechanisms involved in this phenomenon, the largest

particle in this study was selected for further analysis. With a diameter of 0.2 μm, it is approximately

one order of magnitude smaller than the smallest microparticles used in CS techniques. However, it

falls between the sizes of those particles and the largest particles considered in previous works that

reported MD simulations. In contrast with those studies, where jetting was not as pronounced as it is

observed in experiments and FEM simulations, the impact of the 0.2 μm particle in this work

displayed an evident jetting.

A number of MD simulations with different impact velocities were performed. Fig. 7(a) shows the

deformed shape of the 0.2 μm particle at 200 ps after impacting to a substrate at 500, 600, 700, 800,

and 1000 m/s. At 500 m/s, even if the particle underwent significant plastic deformation at the bottom,

it does not exhibit jetting formation. The first sign of particle jetting can be observed for the impacted

particle at 600 m/s, and as the impact velocity increases, the jetting becomes more apparent.

Page 20 of 45
Figure 7 a) The final shape of a 0.2 μm particle at 200 ps after impacting to a substrate at 500, 600,

700, 800, and 1000 m/s, and b) The intensity of the shock wave at 5 ps after impact for each

particle.

The intensity of the shock wave generated after the impact was calculated for each velocity at 5 ps,

and plotted in Fig. 7(b). The intensity of shock waves was obtained by finding the pressure jump on

the pressure distribution along the impact direction. The strength of shock waves decreases as they

propagate [68], as some of their energy is dissipated as heat and is expended by displacing the atoms

along their path. As one should expect, the shock wave’s intensity increases as the impact velocity

rises. The generated shock wave for all cases are a multiple of the HEL for copper, which is between

1 to 3 GPa [69,70]. Therefore, the existence of shock wave greater than the HEL is not the only

condition for the formation of jetting. The shock wave intensity for the 500 m/s is around 8 GPa, and

Page 21 of 45
it reaches to 10 GPa for the 600 m/s. Jetting increases with the intensity of the shock wave, and a

huge amount of jetting occurred at 18 GPa for the 1000 m/s particle.

In solids, shock waves can be categorized as weak or strong shocks with respect to their intensities

[71]. In weak shock regime, plastic deformation occurs behind the shock wave [71,72], with velocities

slower than the elastic precursor, which travels at the sound velocity [73]. The velocity of plastic

wave increases by increasing the impact velocity (or shock wave intensity). Strong shock occurs when

the plastic deformation velocity reaches or exceeds the speed of sound, meaning the plastic

deformation happens ahead of the elastic wave [71,74]. In weak shocks, heterogeneous dislocation

nucleation is dominant, while in strong shocks, the homogeneous dislocation nucleation governs the

plasticity in the material [75]. Therefore, it is very important to understand that the behavior of a

metal subjected to a weak shock is completely different from the one subjected to a strong shock. The

intensity of all shock waves shown in Fig. 7(b) appears to be in the weak shock regime, which for

copper is below 23.4 GPa [19,71]. Hence, at an impact velocity between 600 and 1000 m/s, which is

a common velocity in CS, jetting happens in the weak shock regime: jetting in CS is a consequence

of accumulation of plastic deformation, which occurs heterogeneously behind the shock wave, and is

not a sudden phenomenon that happens at the shock wave.

Fig. 8 shows snapshots of the 0.2 μm particle deformation along with the pressure distributions after

impacting at 500 and 700 m/s. The maximum localized pressure (or hydrostatic pressure), 𝑃𝑚𝑎𝑥 , that

the 500 m/s particle experienced during the impact was slightly larger than the one that occurred for

the 700 m/s particle, i.e., 28 GPa at 500 m/s versus 25 GPa at 700 m/s. The higher pressure for the

500 m/s occurred because the compressive load (or force) between the atomic layers were applied on

a smaller area (the contact area is smaller than the one for the 700 m/s particle). In addition, the time

period in which this maximum pressure applied is different for each impact velocity. At 500 m/s, the

𝑃𝑚𝑎𝑥 happened around 35 ps, while at 700 m/s particle, it was observed at around 15 ps.

At the early stage of the impact, a high pressure is formed in both particles at the contact zone, with

its maximum at the impact center. The formation and interaction of tensile waves with the shock wave

Page 22 of 45
can be seen. At 15 ps, the effect of tensile waves interaction with the shock wave is apparent at the

particle bottom, see “tensile wave propagation area” in Fig. 8.

Large pressure gradients are formed near the contact edges of particles at around 30 and 35 ps, which

is due to the propagation of tensile waves. However, the contact pressure is still extremely high and

thus, keep the region in the compressive state for both particles. At around 35 ps, the shock wave

reaches the particles’ top and start to reflect as tensile waves towards the particles’ bottom.

Tensile waves are the predominant waves in the top part of particles at around 50 ps, while the bottom

part is mostly under compression. In both particles, two extensive tensile waves can be seen at the

top of the particles, which are moving towards each other, shown by blue dotted line on the 50 ps

snapshot in Fig. 8 (due to the symmetry of waves’ propagation, only tensile waves in the left side of

particles are highlighted). At 50 ps, for the impact velocity of 500 m/s the atoms that just met the

tensile waves are still in the compressive zone (around 1 GPa), while those atoms when impact took

place at 700 m/s are in the tension state by around 1 GPa, which is related to the stronger generated

shock wave at 700 m/s.

Between 50 and 55 ps, the tensile waves met each other in both particles, see the enlargement of

selected areas at 500 m/s. The effect of this collision is more pronounced for the impact at 700 m/s,

as a significant decrease (around 5 GPa) in the compression state of the particle bottom occurred.

After the collision, the tensile waves continue forwarding to the particles’ free surfaces and contact

edges. At 65 ps, the widening of the bottom part of the particle that impacted to the substrate at 700

m/s is more noticeable.

Page 23 of 45
Figure 8 Pressure distribution of a particle during the deformation after impact at 700 m/s. The

green dotted line on the 5 and 15 ps snapshots indicates the shock wave inside the particles. The

blue dotted line illustrates tensile waves. It should be noted that due to the symmetry of waves’

propagation, only tensile waves in the left side of particles are shown by the blue dotted line. In

addition, the affected area with the tensile waves is also demonstrated on the 15 ps snapshots.

The higher particle kinetic energy resulted in a stronger shock wave and larger contact area between

the particle and the substrate, at the early stage of the impact. In Fig. 8, it is shown that the particle

that impacted the substrate at 500 m/s particle experienced a higher localized pressure (hydrostatic

pressure), whereas its total deformation was less. Therefore, the high-pressure gradient was not the

main factor of plastic deformation of particles. The main factor appears to be the intensity of the

shock wave, which has a direct relationship with the deformation rate and the activation of plasticity

mechanisms, which will be discussed in the next section.

The effects of shock wave on the plasticity behavior

Page 24 of 45
The glide of Shockley partial dislocations with Burgers vector 1/6 <112> is the dominant plasticity

mechanism in FCC metals with low and medium stacking faults energies (like copper) [76]. In a

crystalline material subjected to strain, regions of concentrated glides, and high dislocation

multiplication rates create a banded structure called “slip bands” [77]. Slip bands are one of many

types of banded structure, like micro bands, shear bands, etc. Figs. 9 and 10 display the atomic strain

snapshots of particles during deformation. The atomic strain (shear strain) analysis has been

extensively used in the literature to characterize plasticity mechanisms such as slip bands, shear bands,

and twining during the deformation of crystalline and non-crystalline materials [78–81].

Fig. 9 displays the pressure distribution and the atomic shear strain of the 0.2 μm particle after the

impact at 700 m/s. As mentioned in the previous section, the generated shock wave at 700 m/s was in

the weak shock regime. Consequently, plastic deformation occurred behind the shock wave. In Fig.

9, it can be seen that although the shock wave was already generated at 5 ps, plastic deformation had

not been developed yet at this time. At 10 ps, formation of a slip band like feature at the bottom of

the particle can be observed. The nucleation happened in areas which are shown as “Tensile wave

propagation area”, see Figs. 9(a-c) at 10 ps. Tensile waves created a shear stress along their path,

causing heterogeneously nucleation of dislocations at the particle bottom. At 15 ps, slip bands glided

towards the particle’s center, while new slip bands were also generated. By looking at Fig. 9, it is

apparent that the hydrostatic pressure at the particle’s south pole did not cause any plastic deformation,

though its intensity reached to 25 GPa.

Page 25 of 45
Figure 9 a) Pressure distribution, b) Atomic shear strain of the 0.2 μm particle at 5, 10, and 15 ps

after impacting at 700 m/s, and c) Enlargement of selected area in (b).

In Fig. 10(a), the formation and evolution of slip bands during the particle deformation is shown. At

20 ps, there are some interactions between some of the bands, but the density of the bands is still

relatively low. At 35 ps, the generation of first bands near the particle’s contact edges can be observed.

By continuing the deformation, the density of the slip bands increases. In Fig. 8, it was shown that

the shock wave reached to the particle top, and reflected back towards the particle bottom as a tensile

wave after 35 ps. Afterwards, at 50 ps, the deformation became localized more around the contact

edges, and the width of bands in these areas also became thicker. At 65 ps, the shear localization at

the edges exceeds its stability limit, and the instability begins after this time, as can be seen at 85 and

200 ps, in which the material starts to flow with a fluid-like behavior at the edges, resulting in the

material jetting at this region.

Page 26 of 45
Figure 10 Atomic shear strain in snapshots of the 0.2 μm particle after impacting to a substrate at a)

700 m/s, and b) 500 m/s.

In comparison to the impact at 700 m/s, the nucleation of slip bands during the deformation of the

particle that hit the substrate at 500 m/s occurred homogeneously above the contact surface at a later

time, i.e., around 40 ps, see Fig. 10(b). This homogeneous nucleation of dislocations was due to the

superposition of reflected tensile waves which met at the middle of the particle, see Fig. 8. After the

nucleation of the first bands, the dislocation activity increased, and the slip band network expanded

in all directions, as one can see at 45 and 50 ps. The slip bands’ junctions acted as dislocation sources

and facilitated the plastic deformation in that region. The plastic deformation reached the particle’s

bottom surface and the contact edges at 65 ps, However, in a clear contrast with the impact at 700

m/s, the shear localization was not at the contact edges, but at the regions closer to the contact center,

as can be observed at 80 to 200 ps. As a consequence of this different deformation distribution inside

Page 27 of 45
the particle, the impact velocity of 500 m/s did not display a pronounced jetting zone although the

particle bottom was significantly flattened.

Fig. 11(a) shows the temperature distribution inside the 0.2 μm particle after impacting the substrate

at 700 m/s. By looking at Fig. 10(a) and Fig. 11(a) side by side, the effects of the slip bands’

propagation during the deformation becomes clear: by the formation and gliding of the slip bands,

the material temperature noticeably increased. This temperature rise was more substantial where the

bands encountered each other, as one can observe in the snapshots taken at 20 and 35 ps. As the slip

bands density and their interactions increases the temperature also rises. At 65 ps, when the shear

localization exceeded the material threshold, the temperature of the jetting zone rose significantly to

about 1150 K, while the maximum temperature at the particle center was around 700 K. The high

temperature at the jetting zone promotes the thermal softening, facilitating severe plastic deformation

in this region. It is worth pointing out that, even after elapsed 200 ps, the upper half of the particle

remained near the initial temperature of 300 K. This implies that heat transfer, which in MD is solely

due to phonon transport [82], was not fast enough to raise the temperature of the whole particle despite

the high degree of deformation.

Page 28 of 45
Figure 11 Temperature distribution of the 0.2 μm particle during the impact at a) 700 m/s, and b) at

500 m/s. c) Comparison between the von Mises and the shear stress pattern during the impact.

The temperature distribution for the particle impacting the substrate at 500 m/s is displayed in Fig.

11(b). The material temperature also raises up as the deformation proceeds. For each snapshot, the

maximum temperature was found where the density of the slip bands is maximum. At 80 ps, the

temperature peak was 725 K and was found at the particle center and the contact edges. The

temperature at the contact edges increased to 830 K at 100 ps, but it is still much lower than the

maximum temperature experienced when the 0.2 μm particle collided at 700 m/s. Comparison

between Fig. 11(a) and Fig. 11(b) indicates that temperature plays an essential role in the formation

Page 29 of 45
of jetting, and instability in the material. The temperature effects on the jetting will be discussed more

in the next section.

Examination of Figs. 10 and 11 indicates that although the plasticity in particles began differently,

the overall shear strain pattern was similar for both particles. In [32], it was shown that the shear

strain pattern is independent from the microstructure of the particles during high velocity impacts.

Therefore, the shear strain pattern is independent from both microstructure and the involved plasticity

mechanisms, and is governed mainly by the applied shear stress which has an arrow-head shape.

Regions with the highest shear strain (or temperature) were found near the contact edges and close to

the particle center. These regions experienced the highest applied shear stresses during the impact.

The von Mises and the shear stress pattern during the impact are shown schematically in Fig. 11(c).

The maximum von Mises is obtained at the impact center, and it gradually decreases by receding to

the contact edges and the center of the particle. However, the applied shear stress shows an opposite

pattern compared to the von Mises distribution. The shear stress is almost zero at the impact center,

and it is increasing by receding to the contact edges and the particle center. At the impact center, the

matter is restricted in all directions and the net flow in this region is almost zero. This condition is

due to the large hydrostatic pressure which is formed at this region. By going farther from the impact

center, the hydrostatic pressure decreases, allowing portions of the particle material to undergo shear.

This definition is in agreement with the experimental observations that have been reported in the

literature [7,60,83,84], as the dynamic recrystallization occurs largely at the contact zone, and the

impact center region experienced little to no dynamic recrystallization.

Recently, the “hydrodynamic pressure release” hypothesis has been proposed [16,85–87] as the

mechanism controlling the onset of jetting in CS. This hypothesis suggests that when the high-

pressure wave reaches the edge of the flattened particle, it produces a significant localized tension,

resulting to spall/fragmentation in the form of a jet. However, the current results showed that the

amplitude of the tensile waves reflected from the free surfaces were not large enough to be able to

change the material state from the compression to tension upon the impact, and they did not have a

Page 30 of 45
direct effect on the generation of jetting in CS. As such, the results seem to indicate that the formation

of jetting is due to the localization of shear, following by a significant increase in the material

temperature at the contact edges. This leads to create an instability state and a fluid-like behavior for

the material in that region.

In addition, the so-called “spall process” is an internal failure process, which occurs when the

superposition of reflected tensile waves excess the tensile strength of the material, resulting in rupture

or failure within the body [72,88–90]. It is important to note that spall event cannot occur at the free

surfaces [90]. Thus, the explanation reported in [16] which stated that spall is responsible for jetting

in CS contradicts the spall theory definition [72,88–91].

The role of temperature on jetting

Fig. 12(a) displays the temperature distribution of the 0.2 μm particle at 200 ps after impacting with

different velocities. The maximum temperature found in the jetting area raised by increasing the

impact velocity. However, the increase in the temperature was localized at the contact zone and the

jetting area. At 600 m/s, the minimum temperature of the jetting was 850 K, and its maximum

captured as 1100 K. At 700 m/s which the jetting is more noticeable, an overall increase of the

temperature at the jetting area can be observed, see Fig. 12(a). At that velocity, the average

temperature of the jetting reached to 1200 K. By increasing the impact velocity to 800 m/s, the

maximum temperature (i.e., around 1700 K) in some areas of jetting exceeded the melting point of

copper, resulting fragmentation and material ejecta at the contact edges. At 1000 m/s, the portion of

molten material at the jetting area raised significantly, resulting a more pronounced jetting as can be

seen in Fig. 12(a).

Page 31 of 45
Figure 12 a) Temperature distribution, and b) Microstructure of the 0.2 μm particle for different

impact velocities at 200 ps (Different colors indicate different grains with low and high

misorientation angle). c) enlargement of selected areas in (b), and d) redrawn of (c) by elimination

of atoms that have amorphous structure.

Fig. 12(b) shows the microstructure of the particles at 200 ps after the impact using the Polyhedral

Template Matching (PTM) [92] analysis. In a previous study [32], it was shown that the PTM

algorithm is able to illustrate plasticity mechanisms such as formation of LAGBs, HAGBs, and

twinning. The enlargement of the jetting zone is shown in Fig. 12(c). To have a better idea of the

material status at the jetting zone, all non-crystalline atoms were eliminated in Fig. 12(d). At 600 m/s,

the fraction of non-crystalline structure is negligible and the jetting happened in solid state. At 700

m/s, even though the fraction of non-crystalline structure increased, the average temperature was still

below the melting point and the jetting occurred in solid state. By increasing the impact velocity to

Page 32 of 45
800 m/s, the fraction of crystalline structure decreased significantly and the first sign of molten

material can be observed, see Fig. 12(d). In addition, at 800 m/s, the temperature at the contact zone

was also exceeding the melting point and a thin layer of molten material was formed at the interface.

At 1000 m/s, the majority of atoms lost their crystalline structures and thus, an enormous jetting at

this velocity was formed in liquid state.

Fig. 12 shows that the jetting can occur in both solid (i.e., crystalline and non-crystalline) and liquid

state. By increasing the impact velocity, the material jetting increased, resulting in more deformation

and raising the temperature in that area. Localization of deformation means an increase in the

dislocation density, following by occurrence of solid-state amorphization (700 m/s particle). The

increase in the plastic deformation continues by increasing the impact velocity, resulting the material

temperature exceeded the melting point and the formation of liquid-state at the jetting (800 m/s and

1000 m/s particles).

Conclusion

The objective of this study was to investigate the nature of jetting and reveal the deformation behavior

of this phenomenon at the atomic level. First, the spherical and the quasi-2D models were compared.

It was shown that the same plasticity mechanisms are involved during the deformation of both models.

Second, particles with 40, 80, 120, and 200 nm in diameter were modeled, and the effects of particle

size on the deformation behavior were studied. It was indicated that the fraction of non-crystalline

structure decreases by increasing the particle size. Then, the 0.2 μm particle was selected to explore

the phenomena involved during the formation of jetting in CS. The mechanical responses, crystalline

structure, and microstructural evolution during high velocity impacts were characterized. The primary

findings of this study are as follows:

 The shock wave propagation inside the particles during high velocity impacts were

investigated, and it was demonstrated that the generated shock wave in CS particles is in the

weak shock regime, which means the plastic deformation occurs behind the shock wave.

Page 33 of 45
 It was revealed that expansion waves (or tensile waves) cause the heterogenous nucleation of

dislocations at the particle bottom.

 It was shown that the localization of a band-like structure at the contact edges increased the

material temperature, resulting in thermal softening and raising the deformation at the contact

edges. Thermal softening alongside the localization of shear strain formed an instability at the

contact edges which is the main reason for the formation of jetting.

 It was shown that, unlike the droplet impact and shaped charges processes, the jetting in CS

is not a sudden phenomenon related to “hydrodynamic pressure release”, but is a progressive

phenomenon.

 Microstructural analysis revealed that the jetting can be formed in solid-state crystalline

structure, solid-state amorphous structure, and liquid structure (molten state).

 It was established that due to the extremely high applied shear stress during the impact, the

deformation pattern is independent from the involved plasticity mechanisms.

References

[1] S. Yin, P. Cavaliere, B. Aldwell, R. Jenkins, H. Liao, W. Li, R. Lupoi, Cold spray additive

manufacturing and repair: Fundamentals and applications, Addit. Manuf. 21 (2018) 628–650.

https://doi.org/https://doi.org/10.1016/j.addma.2018.04.017.

[2] G. Prashar, H. Vasudev, A comprehensive review on sustainable cold spray additive

manufacturing: State of the art, challenges and future challenges, J. Clean. Prod. 310 (2021)

127606. https://doi.org/https://doi.org/10.1016/j.jclepro.2021.127606.

[3] A.A. Tiamiyu, Y. Sun, K.A. Nelson, C.A. Schuh, Site-specific study of jetting, bonding, and

local deformation during high-velocity metallic microparticle impact, Acta Mater. 202 (2021)

159–169. https://doi.org/https://doi.org/10.1016/j.actamat.2020.10.057.

Page 34 of 45
[4] K. Kim, M. Watanabe, S. Kuroda, Jetting-Out Phenomenon Associated with Bonding of

Warm-Sprayed Titanium Particles onto Steel Substrate, J. Therm. Spray Technol. 18 (2009)

490. https://doi.org/10.1007/s11666-009-9379-1.

[5] M. Hassani-Gangaraj, D. Veysset, K.A. Nelson, C.A. Schuh, In-situ observations of single

micro-particle impact bonding, Scr. Mater. 145 (2018) 9–13.

https://doi.org/10.1016/J.SCRIPTAMAT.2017.09.042.

[6] A. Viscusi, A. Astarita, R. Della Gatta, F. Rubino, A perspective review on the bonding

mechanisms in cold gas dynamic spray, Surf. Eng. 35 (2019) 743–771.

https://doi.org/10.1080/02670844.2018.1551768.

[7] M. Walker, Microstructure and bonding mechanisms in cold spray coatings, Mater. Sci.

Technol. 34 (2018) 2057–2077. https://doi.org/10.1080/02670836.2018.1475444.

[8] M. Grujicic, C.L. Zhao, W.S. DeRosset, D. Helfritch, Adiabatic shear instability based

mechanism for particles/substrate bonding in the cold-gas dynamic-spray process, Mater.

Des. 25 (2004) 681–688. https://doi.org/http://dx.doi.org/10.1016/j.matdes.2004.03.008.

[9] Y. Sun, D. Veysset, K.A. Nelson, C.A. Schuh, The transition from rebound to bonding in

high-velocity metallic microparticle impacts: jetting-associated power-law divergence, J.

Appl. Mech. (2020) 1–28. https://doi.org/10.1115/1.4047206.

[10] W.-Y. Li, C.-J. Li, H. Liao, Significant influence of particle surface oxidation on deposition

efficiency, interface microstructure and adhesive strength of cold-sprayed copper coatings,

Appl. Surf. Sci. 256 (2010) 4953–4958.

https://doi.org/https://doi.org/10.1016/j.apsusc.2010.03.008.

[11] W.-Y. Li, C.-J. Li, G.-J. Yang, Effect of impact-induced melting on interface microstructure

and bonding of cold-sprayed zinc coating, Appl. Surf. Sci. 257 (2010) 1516–1523.

https://doi.org/10.1016/J.APSUSC.2010.08.089.

Page 35 of 45
[12] C.-J. Li, W.-Y. Li, Y.-Y. Wang, Formation of metastable phases in cold-sprayed soft metallic

deposit, Surf. Coatings Technol. 198 (2005) 469–473.

https://doi.org/https://doi.org/10.1016/j.surfcoat.2004.10.063.

[13] Y. Xiong, X. Xiong, S. Yoon, G. Bae, C. Lee, Dependence of Bonding Mechanisms of Cold

Sprayed Coatings on Strain-Rate-Induced Non-Equilibrium Phase Transformation, J. Therm.

Spray Technol. 20 (2011) 860–865. https://doi.org/10.1007/s11666-011-9634-0.

[14] P.C. King, C. Busch, T. Kittel-Sherri, M. Jahedi, S. Gulizia, Interface melding in cold spray

titanium particle impact, Surf. Coatings Technol. 239 (2014) 191–199.

https://doi.org/10.1016/J.SURFCOAT.2013.11.039.

[15] P.C. King, S.H. Zahiri, M. Jahedi, Microstructural Refinement within a Cold-Sprayed

Copper Particle, Metall. Mater. Trans. A. 40 (2009) 2115–2123.

https://doi.org/10.1007/s11661-009-9882-5.

[16] M. Hassani-Gangaraj, D. Veysset, V.K. Champagne, K.A. Nelson, C.A. Schuh, Adiabatic

shear instability is not necessary for adhesion in cold spray, Acta Mater. 158 (2018) 430–

439. https://doi.org/10.1016/J.ACTAMAT.2018.07.065.

[17] J.A. Zukas, ed., Chapter 3 Shock waves in solids, in: Introd. to Hydrocodes, Elsevier, 2004:

pp. 75–102. https://doi.org/https://doi.org/10.1016/S0922-5382(04)80004-7.

[18] P. Peralta, E. Loomis, Y. Chen, A. Brown, R. McDonald, K. Krishnan, H. Lim, Grain

orientation effects on dynamic strength of FCC multicrystals at low shock pressures: a

hydrodynamic instability study, Philos. Mag. Lett. 95 (2015) 67–76.

https://doi.org/10.1080/09500839.2014.994571.

[19] N.K. Bourne, The threshold for hydrodynamic behaviour in solids under extreme

compression, J. Appl. Phys. 116 (2014) 93505. https://doi.org/10.1063/1.4894138.

[20] H.-S. Park, B.A. Remington, R.C. Becker, J. V Bernier, R.M. Cavallo, K.T. Lorenz, S.M.

Page 36 of 45
Pollaine, S.T. Prisbrey, R.E. Rudd, N.R. Barton, Strong stabilization of the Rayleigh–Taylor

instability by material strength at megabar pressures, Phys. Plasmas. 17 (2010) 56314.

https://doi.org/10.1063/1.3363170.

[21] A.R. Piriz, J.J. López Cela, N.A. Tahir, D.H.H. Hoffmann, Richtmyer-Meshkov instability in

elastic-plastic media, Phys. Rev. E. 78 (2008) 56401.

https://doi.org/10.1103/PhysRevE.78.056401.

[22] T.J. Vogler, L.C. Chhabildas, Strength behavior of materials at high pressures, Int. J. Impact

Eng. 33 (2006) 812–825. https://doi.org/https://doi.org/10.1016/j.ijimpeng.2006.09.069.

[23] W.T. Buttler, D.M. Oró, D.L. Preston, K.O. Mikaelian, F.J. Cherne, R.S. Hixson, F.G.

Mariam, C. Morris, J.B. Stone, G. Terrones, D. Tupa, Unstable Richtmyer–Meshkov growth

of solid and liquid metals in vacuum, J. Fluid Mech. 703 (2012) 60–84.

https://doi.org/10.1017/jfm.2012.190.

[24] E. Zasimchuk, T. Turchak, N. Chausov, Hydrodynamic plastic flow in metal materials,

Results Mater. 6 (2020) 100090. https://doi.org/https://doi.org/10.1016/j.rinma.2020.100090.

[25] Y.G. Gordienko, E.E. Zasimchuk, Synergetic model of structure formation during plastic

deformation of crystals, Philos. Mag. A. 70 (1994) 99–107.

https://doi.org/10.1080/01418619408242540.

[26] E. Zasimchuk, L. Markashova, O. Baskova, T. Turchak, N. Chausov, V. Hutsaylyuk, V.

Berezin, Influence of Combined Loading on Microstructure and Properties of Aluminum

Alloy 2024-T3, J. Mater. Eng. Perform. 22 (2013) 3421–3429.

https://doi.org/10.1007/s11665-013-0630-z.

[27] E. Zasimchuk, O. Baskova, O. Gatsenko, T. Turchak, Universal Mechanism of Viscoplastic

Deformation of Metallic Materials Far from Thermodynamics Equilibrium, J. Mater. Eng.

Perform. 27 (2018) 4183–4196. https://doi.org/10.1007/s11665-018-3515-3.

Page 37 of 45
[28] P. Gao, C. Zhang, R. Wang, G. Deng, J. Li, L. Su, Tamping effect during additive

manufacturing of copper coating by cold spray: A comprehensive molecular dynamics study,

Addit. Manuf. 66 (2023) 103448.

https://doi.org/https://doi.org/10.1016/j.addma.2023.103448.

[29] Q. Jiang, H. Liu, J. Li, D. Yang, Y. Zhang, W. Yang, Atomic-level understanding of

crystallization in the selective laser melting of Fe50Ni50 amorphous alloy, Addit. Manuf. 34

(2020) 101369. https://doi.org/https://doi.org/10.1016/j.addma.2020.101369.

[30] F. Wang, S. You, D. Jiang, F. Ning, Study on sintering mechanism for extrusion-based

additive manufacturing of stainless steel through molecular dynamics simulation, Addit.

Manuf. 58 (2022) 102991. https://doi.org/https://doi.org/10.1016/j.addma.2022.102991.

[31] S. Rahmati, A. Zúñiga, B. Jodoin, R.G.A. Veiga, Deformation of copper particles upon

impact: A molecular dynamics study of cold spray, Comput. Mater. Sci. 171 (2020).

https://doi.org/10.1016/j.commatsci.2019.109219.

[32] S. Rahmati, R.G.A. Veiga, B. Jodoin, A. Zúñiga, Crystal orientation and grain boundary

effects on plastic deformation of FCC particles under high velocity impacts, Materialia. 15

(2021) 101004. https://doi.org/https://doi.org/10.1016/j.mtla.2021.101004.

[33] S. Rahmati, B. Jodoin, Physically Based Finite Element Modeling Method to Predict

Metallic Bonding in Cold Spray, J. Therm. Spray Technol. 29 (2020) 611–629.

https://doi.org/10.1007/s11666-020-01000-1.

[34] S. Rahmati, R.G.A. Veiga, A. Zúñiga, B. Jodoin, A Numerical Approach to Study the Oxide

Layer Effect on Adhesion in Cold Spray, J. Therm. Spray Technol. (2021).

https://doi.org/10.1007/s11666-021-01245-4.

[35] V. Yamakov, D. Wolf, S.R. Phillpot, A.K. Mukherjee, H. Gleiter, Dislocation processes in

the deformation of nanocrystalline aluminium by molecular-dynamics simulation, Nat.

Mater. 1 (2002) 45–49. https://doi.org/10.1038/nmat700.

Page 38 of 45
[36] O. Waseda, H. Goldenstein, G.F.B.L. e Silva, A. Neiva, P. Chantrenne, J. Morthomas, M.

Perez, C.S. Becquart, R.G.A. Veiga, Stability of nanocrystalline Ni-based alloys: coupling

Monte Carlo and molecular dynamics simulations, Model. Simul. Mater. Sci. Eng. 25 (2017)

75005. https://doi.org/10.1088/1361-651x/aa83ef.

[37] S. Kurian, R. Mirzaeifar, Selective laser melting of aluminum nano-powder particles, a

molecular dynamics study, Addit. Manuf. 35 (2020) 101272.

https://doi.org/https://doi.org/10.1016/j.addma.2020.101272.

[38] B.C. Johnson, T.J. Bowling, H.J. Melosh, Jetting during vertical impacts of spherical

projectiles, Icarus. 238 (2014) 13–22.

https://doi.org/https://doi.org/10.1016/j.icarus.2014.05.003.

[39] K.K. Haller, Y. Ventikos, D. Poulikakos, P. Monkewitz, Computational study of high-speed

liquid droplet impact, J. Appl. Phys. 92 (2002) 2821–2828.

https://doi.org/10.1063/1.1495533.

[40] Z. Shen, R.H. Wagoner, W.A.T. Clark, Dislocation and grain boundary interactions in

metals, Acta Metall. 36 (1988) 3231–3242. https://doi.org/https://doi.org/10.1016/0001-

6160(88)90058-2.

[41] K. Wen, X.-W. Chen, Failure evolution in hypervelocity impact of Al spheres onto thin Al

plates, Int. J. Impact Eng. 147 (2021) 103727.

https://doi.org/https://doi.org/10.1016/j.ijimpeng.2020.103727.

[42] K. Kurosawa, T. Okamoto, H. Genda, Hydrocode modeling of the spallation process during

hypervelocity impacts: Implications for the ejection of Martian meteorites, Icarus. 301 (2018)

219–234. https://doi.org/https://doi.org/10.1016/j.icarus.2017.09.015.

[43] K. Wen, X.-W. Chen, D.-N. Di, Modeling on the shock wave in spheres hypervelocity

impact on flat plates, Def. Technol. 15 (2019) 457–466.

https://doi.org/https://doi.org/10.1016/j.dt.2019.01.006.

Page 39 of 45
[44] B. Li, F.P. Zhao, H.A. Wu, S.N. Luo, Microstructure effects on shock-induced surface

jetting, J. Appl. Phys. 115 (2014) 73504. https://doi.org/10.1063/1.4865798.

[45] B. Daneshian, F. Gaertner, H. Assadi, D. Hoeche, W. Weber, T. Klassen, Size Effects of

Brittle Particles in Aerosol Deposition—Molecular Dynamics Simulation, J. Therm. Spray

Technol. 30 (2021) 503–522. https://doi.org/10.1007/s11666-020-01149-9.

[46] C.A. Duarte, C. Li, B.W. Hamilton, A. Strachan, M. Koslowski, Continuum and molecular

dynamics simulations of pore collapse in shocked β-tetramethylene tetranitramine (β-HMX)

single crystals, J. Appl. Phys. 129 (2021) 15904. https://doi.org/10.1063/5.0025050.

[47] J.E. Field, M.B. Lesser, J.P. Dear, D. Tabor, Studies of two-dimensional liquid-wedge impact

and their relevance to liquid-drop impact problems, Proc. R. Soc. London. A. Math. Phys.

Sci. 401 (1985) 225–249. https://doi.org/10.1098/rspa.1985.0096.

[48] M. Plimpton, Steve and Pollock, Roy and Stevens, Particle-Mesh Ewald and rRESPA for

Parallel Molecular Dynamics Simulations, PPSC. (1997).

[49] Y. Mishin, M.J. Mehl, D.A. Papaconstantopoulos, A.F. Voter, J.D. Kress, Structural stability

and lattice defects in copper: Ab initio, tight-binding, and embedded-atom calculations, Phys.

Rev. B. 63 (2001) 224106. https://doi.org/10.1103/PhysRevB.63.224106.

[50] L.M. Pereira, S. Rahmati, A. Zúñiga, B. Jodoin, R.G.A. Veiga, Atomistic study of

metallurgical bonding upon the high velocity impact of fcc core-shell particles, Comput.

Mater. Sci. 186 (2021) 110045.

https://doi.org/https://doi.org/10.1016/j.commatsci.2020.110045.

[51] M. Razavipour, S. Rahmati, A. Zúñiga, D. Criado, B. Jodoin, Bonding Mechanisms in Cold

Spray: Influence of Surface Oxidation During Powder Storage, J. Therm. Spray Technol.

(2020). https://doi.org/10.1007/s11666-020-01123-5.

[52] D. Veysset, A.J. Hsieh, S. Kooi, A.A. Maznev, K.A. Masser, K.A. Nelson, Dynamics of

Page 40 of 45
supersonic microparticle impact on elastomers revealed by real–time multi–frame imaging,

Sci. Rep. 6 (2016) 25577. https://doi.org/10.1038/srep25577.

[53] W. Xie, A. Alizadeh-Dehkharghani, Q. Chen, V.K. Champagne, X. Wang, A.T. Nardi, S.

Kooi, S. Müftü, J.-H. Lee, Dynamics and extreme plasticity of metallic microparticles in

supersonic collisions, Sci. Rep. 7 (2017) 5073. https://doi.org/10.1038/s41598-017-05104-7.

[54] S. Guetta, M.H. Berger, F. Borit, V. Guipont, M. Jeandin, M. Boustie, Y. Ichikawa, K.

Sakaguchi, K. Ogawa, Influence of particle velocity on adhesion of cold-sprayed splats, J.

Therm. Spray Technol. 18 (2009). https://doi.org/10.1007/s11666-009-9327-0.

[55] Y. Xiong, K. Kang, G. Bae, S. Yoon, C. Lee, Dynamic amorphization and recrystallization of

metals in kinetic spray process, Appl. Phys. Lett. 92 (2008) 194101.

https://doi.org/10.1063/1.2928218.

[56] Q. Wang, D. Qiu, Y. Xiong, N. Birbilis, M.-X. Zhang, High resolution microstructure

characterization of the interface between cold sprayed Al coating and Mg alloy substrate,

Appl. Surf. Sci. 289 (2014) 366–369.

https://doi.org/https://doi.org/10.1016/j.apsusc.2013.10.168.

[57] K.H. Ko, J.O. Choi, H. Lee, Y.K. Seo, S.P. Jung, S.S. Yu, Cold spray induced amorphization

at the interface between Fe coatings and Al substrate, Mater. Lett. 149 (2015) 40–42.

https://doi.org/https://doi.org/10.1016/j.matlet.2015.02.118.

[58] T. Liu, J.D. Leazer, L.N. Brewer, Particle deformation and microstructure evolution during

cold spray of individual Al-Cu alloy powder particles, Acta Mater. 168 (2019) 13–23.

https://doi.org/https://doi.org/10.1016/j.actamat.2019.01.054.

[59] Y. Zou, W. Qin, E. Irissou, J.-G. Legoux, S. Yue, J.A. Szpunar, Dynamic recrystallization in

the particle/particle interfacial region of cold-sprayed nickel coating: Electron backscatter

diffraction characterization, Scr. Mater. 61 (2009) 899–902.

https://doi.org/10.1016/J.SCRIPTAMAT.2009.07.020.

Page 41 of 45
[60] R. Nikbakht, M. Saadati, H. Assadi, K. Jahani, B. Jodoin, Dynamic microstructure evolution

in cold sprayed NiTi composite coatings, Surf. Coatings Technol. 421 (2021) 127456.

https://doi.org/https://doi.org/10.1016/j.surfcoat.2021.127456.

[61] F. Findik, Recent developments in explosive welding, Mater. Des. 32 (2011) 1081–1093.

https://doi.org/https://doi.org/10.1016/j.matdes.2010.10.017.

[62] M. Rein, Phenomena of liquid drop impact on solid and liquid surfaces, Fluid Dyn. Res. 12

(1993) 61–93. https://doi.org/10.1016/0169-5983(93)90106-k.

[63] L.E. Murr, C.-S. Niou, E.P. Garcia, E. Ferreyra, T.J.M. Rivas, J.C. Sanchez, Comparison of

jetting-related microstructures associated with hypervelocity impact crater formation in

copper targets and copper shaped charges, Mater. Sci. Eng. A. 222 (1997) 118–132.

https://doi.org/https://doi.org/10.1016/S0921-5093(96)10518-9.

[64] Z.X. Zhang, Rock Fracture and Blasting: Theory and Applications, Elsevier Science, 2016.

https://books.google.ca/books?id=ILT4CQAAQBAJ.

[65] J. Zukas, Introduction to Hydrocodes, Elsevier Science, 2004.

https://books.google.ca/books?id=qEisMAqkut0C.

[66] Y. Waseda, E. Matsubara, K. Shinoda, X-Ray Diffraction Crystallography: Introduction,

Examples and Solved Problems, Springer Berlin Heidelberg, 2011.

https://books.google.ca/books?id=vk9fnLH56DYC.

[67] R.O. Gould, W. Borchardt-Ott, Crystallography: An Introduction, Springer Berlin

Heidelberg, 2011. https://books.google.ca/books?id=DgI6kgEACAAJ.

[68] P. Hough, T.J. Kelly, C. Fallon, C. McLoughlin, P. Hayden, E.T. Kennedy, J.P. Mosnier,

S.S. Harilal, J.T. Costello, Enhanced shock wave detection sensitivity for laser-produced

plasmas in low pressure ambient gases using interferometry, Meas. Sci. Technol. 23 (2012)

125204. https://doi.org/10.1088/0957-0233/23/12/125204.

Page 42 of 45
[69] D. Seif, G. Po, R. Crum, V. Gupta, N.M. Ghoniem, Shock-induced plasticity and the

Hugoniot elastic limit in copper nano films and rods, J. Appl. Phys. 115 (2014) 54301.

https://doi.org/10.1063/1.4863720.

[70] S. V Razorenov, G.I. Kanel, K. Baumung, H.J. Bluhm, Hugoniot Elastic Limit and Spall

Strength of Aluminum and Copper Single Crystals over a Wide Range of Strain Rates and

Temperatures, AIP Conf. Proc. 620 (2002) 503–506. https://doi.org/10.1063/1.1483587.

[71] N.K. Bourne, F.L. Bourne, On the transition from weak to strong shock response, J. Appl.

Phys. 131 (2022) 145901. https://doi.org/10.1063/5.0084553.

[72] N. Bourne, Materials in Mechanical Extremes: Fundamentals and Applications, Cambridge

University Press, 2013. https://books.google.ca/books?id=uXkhAwAAQBAJ.

[73] D.C. Wallace, Flow process of weak shocks in solids, Phys. Rev. B. 22 (1980) 1487–1494.

https://doi.org/10.1103/PhysRevB.22.1487.

[74] C.S. Coffey, J. Sharma, Plastic deformation, energy dissipation, and initiation of crystalline

explosives, Phys. Rev. B. 60 (1999) 9365–9371. https://doi.org/10.1103/PhysRevB.60.9365.

[75] R.A. Austin, D.L. McDowell, A dislocation-based constitutive model for viscoplastic

deformation of fcc metals at very high strain rates, Int. J. Plast. 27 (2011) 1–24.

[76] D. Hull, D.J. Bacon, Introduction to Dislocations, Elsevier Science, 2011.

https://books.google.ca/books?id=MvSvqll6ct0C.

[77] Z.Q. Wang, I.J. Beyerlein, R. LeSar, Slip band formation and mobile dislocation density

generation in high rate deformation of single fcc crystals, Philos. Mag. 88 (2008) 1321–1343.

https://doi.org/10.1080/14786430802129833.

[78] Q.-K. Li, M. Li, Assessing the critical sizes for shear band formation in metallic glasses from

molecular dynamics simulation, Appl. Phys. Lett. 91 (2007) 231905.

https://doi.org/10.1063/1.2821832.

Page 43 of 45
[79] S. Feng, L. Qi, L. Wang, S. Pan, M. Ma, X. Zhang, G. Li, R. Liu, Atomic structure of shear

bands in Cu64Zr36 metallic glasses studied by molecular dynamics simulations, Acta Mater.

95 (2015) 236–243. https://doi.org/https://doi.org/10.1016/j.actamat.2015.05.047.

[80] J. Wang, P.D. Hodgson, J. Zhang, W. Yan, C. Yang, Effects of pores on shear bands in

metallic glasses: A molecular dynamics study, Comput. Mater. Sci. 50 (2010) 211–217.

https://doi.org/https://doi.org/10.1016/j.commatsci.2010.08.001.

[81] F. Di Gioacchino, T.E.J. Edwards, G.N. Wells, W.J. Clegg, A new mechanism of strain

transfer in polycrystals, Sci. Rep. 10 (2020) 10082. https://doi.org/10.1038/s41598-020-

66569-7.

[82] A.J.H. McGaughey, M. Kaviany, Phonon Transport in Molecular Dynamics Simulations:

Formulation and Thermal Conductivity Prediction, in: G.A. Greene, J.P. Hartnett†, A. Bar-

Cohen, Y.I. Cho (Eds.), Elsevier, 2006: pp. 169–255.

https://doi.org/https://doi.org/10.1016/S0065-2717(06)39002-8.

[83] C. Chen, Y. Xie, R. Huang, S. Deng, Z. Ren, H. Liao, On the role of oxide film’s cleaning

effect into the metallurgical bonding during cold spray, Mater. Lett. 210 (2018) 199–202.

https://doi.org/https://doi.org/10.1016/j.matlet.2017.09.024.

[84] A. Nastic, B. Jodoin, J.-G. Legoux, D. Poirier, Particle Impact Characteristics Influence on

Cold Spray Bonding: Investigation of Interfacial Phenomena for Soft Particles on Hard

Substrates, J. Therm. Spray Technol. 30 (2021) 2013–2033. https://doi.org/10.1007/s11666-

021-01272-1.

[85] S. Suresh, S.-W. Lee, M. Aindow, H.D. Brody, V.K. Champagne, A.M. Dongare, Mesoscale

modeling of jet initiation behavior and microstructural evolution during cold spray single

particle impact, Acta Mater. 182 (2020) 197–206.

https://doi.org/https://doi.org/10.1016/j.actamat.2019.10.039.

[86] A. Fardan, C.C. Berndt, R. Ahmed, Numerical modelling of particle impact and residual

Page 44 of 45
stresses in cold sprayed coatings: A review, Surf. Coatings Technol. 409 (2021) 126835.

https://doi.org/https://doi.org/10.1016/j.surfcoat.2021.126835.

[87] J. Lienhard, C. Crook, M.Z. Azar, M. Hassani, D.R. Mumm, D. Veysset, D. Apelian, K.A.

Nelson, V. Champagne, A. Nardi, C.A. Schuh, L. Valdevit, Surface oxide and hydroxide

effects on aluminum microparticle impact bonding, Acta Mater. 197 (2020) 28–39.

https://doi.org/https://doi.org/10.1016/j.actamat.2020.07.011.

[88] D. Grady, Physics of Shock Impact: Fundamentals and Dynamic Failure, Institute of Physics

Publishing, 2017. https://books.google.ca/books?id=eEcrtAEACAAJ.

[89] J.N. Johnson, Dynamic fracture and spallation in ductile solids, J. Appl. Phys. 52 (1981)

2812–2825. https://doi.org/10.1063/1.329011.

[90] D.E. Grady, The spall strength of condensed matter, J. Mech. Phys. Solids. 36 (1988) 353–

384. https://doi.org/https://doi.org/10.1016/0022-5096(88)90015-4.

[91] L. Davison, A.L. Stevens, M.E. Kipp, Theory of spall damage accumulation in ductile

metals, J. Mech. Phys. Solids. 25 (1977) 11–28. https://doi.org/https://doi.org/10.1016/0022-

5096(77)90017-5.

[92] P.M. Larsen, S. Schmidt, J. Schiøtz, Robust structural identification via polyhedral template

matching, Model. Simul. Mater. Sci. Eng. 24 (2016) 55007. https://doi.org/10.1088/0965-

0393/24/5/055007.

Page 45 of 45
Declaration of Interest Statement

Declaration of interests

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

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

You might also like