You are on page 1of 7

Theoretically optimized hybrid magnetic

nanoparticle concentrations for functional


gradient nanocomposites
Cite as: AIP Advances 10, 105209 (2020); https://doi.org/10.1063/5.0023941
Submitted: 03 August 2020 . Accepted: 12 September 2020 . Published Online: 05 October 2020

Deshan Liang, Xiao Cui, Xingqiao Ma, Xiaoming Shi, Jing Wang, Hasnain Mehdi Jafri, Junsheng Wang,
Zhengzhi Wang, and Houbing Huang

COLLECTIONS

Paper published as part of the special topic on Chemical Physics, Energy, Fluids and Plasmas, Materials Science
and Mathematical Physics

ARTICLES YOU MAY BE INTERESTED IN

Phase-field simulation of two-dimensional topological charges in nematic liquid crystals


Journal of Applied Physics 128, 124701 (2020); https://doi.org/10.1063/5.0021079

Electrical properties of carbon nanotube/liquid metal/rubber nanocomposites


AIP Advances 10, 105106 (2020); https://doi.org/10.1063/5.0027021

Research on brushless DC motor control system based on fuzzy parameter adaptive PI


algorithm
AIP Advances 10, 105208 (2020); https://doi.org/10.1063/5.0025000

AIP Advances 10, 105209 (2020); https://doi.org/10.1063/5.0023941 10, 105209

© 2020 Author(s).
AIP Advances ARTICLE scitation.org/journal/adv

Theoretically optimized hybrid magnetic


nanoparticle concentrations for functional
gradient nanocomposites
Cite as: AIP Advances 10, 105209 (2020); doi: 10.1063/5.0023941
Submitted: 3 August 2020 • Accepted: 12 September 2020 •
Published Online: 5 October 2020

Deshan Liang,1 Xiao Cui,1 Xingqiao Ma,1,a) Xiaoming Shi,2 Jing Wang,2 Hasnain Mehdi Jafri,2
Junsheng Wang, Zhengzhi Wang, and Houbing Huang2,b)
2 3

AFFILIATIONS
1
Department of Physics, University of Science and Technology Beijing, Beijing 100083, China
2
Advanced Research Institute of Multidisciplinary Science, Beijing Institute of Technology, Beijing 100081, China
3
Department of Engineering Mechanics, School of Civil Engineering, Wuhan University, Wuhan, Hubei 430072, China

a)
Email: xqma@sas.ustb.edu.cn
b)
Author to whom correspondence should be addressed: hbhuang@bit.edu.cn

ABSTRACT
Magnetically actuated functional gradient nanocomposites have widely been used for ultra-durable biomimetic interfaces and surfaces. How-
ever, the mechanical and thermal mismatches in integrated systems containing dissimilar materials or structures usually cause failures. By
modulating the concentration of magnetic particles, a suitable mechanical gradient morphology can be generated to match different integrated
systems. In this work, a new model is developed to describe magnetic particle motion under the magnetic field. Hybrid nano-reinforcements
with two different magnetic particle sizes and concentrations were employed to optimize the magnetic particle concentration gradient. It was
observed that the diversification of concentration distribution can be achieved by tuning the sizes and concentrations of nanoparticles. The
present study, therefore, contributes toward the understanding of the transport mechanism of magnetic-field-actuated functional gradient
nanocomposites and provides guidance for experiments to design ultra-durable biomimetic interfaces and surfaces.
© 2020 Author(s). All article content, except where otherwise noted, is licensed under a Creative Commons Attribution (CC BY) license
(http://creativecommons.org/licenses/by/4.0/). https://doi.org/10.1063/5.0023941., s

I. INTRODUCTION wearable electronics,17,18 and other fields.19–23 However, problems


exist in practical applications like hardness insufficiency in the coat-
Magnetically actuated functional gradient nanocomposites ing industry24,25 and lower crack penetration in biomedical implants
have been widely applied in ultra-durable biomimetic interfaces and and restorations,22 where most failures are caused by mechani-
surfaces. For example, to improve resistance to contact damage and cal/thermal mismatches in integrated systems containing dissimilar
layer delamination,1–3 the shell of a mollusk undergoes ∼40% grad- materials or structures.
ual changes between inter-crystalline soft organic layers and stiff The concentration gradient of magnetic nanoparticles (NPs)
ceramic crystallites.4 It is a wonderful idea to mimic such processes can be modulated under the external magnetic field, and magnet-
in artificial composites. Fabricating controllable functional gradient ically actuated FGNCs will get suitable mechanical properties to
nanocomposites (FGNCs) is a great achievement in mimicking the enhance their hardness and toughness. Several schemes have been
gradient structure of nature. FGNCs can achieve better mechani- applied to manipulate the nanoparticle distributions, for example,
cal properties5,6 by gradient distribution to satisfy a wide range of using a micro-electro-magnetic matrix and a micro-electro-magnet
industrial demands, including greater stiffness, super strength, and ring trap to control and manipulate magnetic NPs,26 using uniform
fracture resistance.7–12 Therefore, it is widely used in the fields of and non-uniform magnetic fields to control the motion of particles
biomaterials,13–16 multilayer graphite nanocomposites, flexible and to prepare magnetic-field-sensitive gel beads and monolith gels,27

AIP Advances 10, 105209 (2020); doi: 10.1063/5.0023941 10, 105209-1


© Author(s) 2020
AIP Advances ARTICLE scitation.org/journal/adv

and using a self-assembly system with soft-magnetic hollow cylinder the right side Ff (t) is the thermodynamic fluctuating force of the
template elements embedded in a nonmagnetic substrate to achieve magnetic NPs, and the second term Fη (t) is the viscous dragging
extraordinary control of particle motion.28 The directional motion force.
of uniformly small-sized magnetic NPs by a single cylindrical mag- For a magnetic nanoparticle submerged in a viscous fluid and
net has also been reported, but the maximum packing fraction of subjected to a gradient magnetic field, the magnetic force (Fm )
particles was limited.29 applied to the particle can be calculated as35
In this paper, we studied the transport mechanism of differ-
ent sized core–shell particles (Fe3 O4 @SiO2 ). First, we calculated the (χp − χf )Vp (B ⋅ ∇B)
Fm = , (2)
magnetic force and spatial concentration gradient distribution of μ0
NPs (20 nm–100 nm). Later, we adopted hybrid particles (large and
where χ p and χ f are the magnetic susceptibility of the magnetic
small sizes) and simulated the spatial concentration distribution of
nanoparticle and the polymer matrix, respectively, V p is the volume
hybrid NPs. The results demonstrate that hybrid NPs can achieve an
of the magnetic nanoparticle, B = μ0 H is the magnetic flux density
optimum magnetic particle concentration gradient.
of the applied magnetic field, where μ0 is the permeability of the
free space and H is the applied magnetic field at the center of the
II. NUMERICAL METHODS nanoparticle, and ∇ is the gradient operator.
The viscous force Fη (t) can be calculated by
Redistributions of magnetic NPs in a fluid matrix under a
magnetic field can be simulated by the Brownian motion under a dr
Fη (t) = −6πηRp , (3)
magnetic field.30 This motion process is governed by several forces dt
including magnetic force, fluidic viscosity resistance, and thermal where η is the viscous constant of the liquid and Rp is the radius of
fluctuation.31–33 In our model, we take magnetic force, fluidic vis- the nanoparticle. The diffusion constant D can be expressed by the
cosity resistance, and thermodynamic fluctuating force into account. kB T
Einstein relation D = 6πηR .
The transport of NPs can be assumed as a Brownian motion pro- p

cess with the evolving particle positions predicted by Langevin’s In a system without damping, the fluctuating force component
equation,34 of a nanoparticle can be calculated by
d2 r d2 r
m = Ff (t) + Fη (t) + Fm (r), (1) Ff (t) = m , (4)
dt 2 dt 2
where m is the mass of the nanoparticle, r is the position of the par- where x is the displacement, t is the time, and m is the mass of a
ticle within the polymer matrix, and t is the time. The first term of nanoparticle. Besides,

FIG. 1. (a) Schematics of two magnets,


the nanopillar, and different sized NPs
and the right side is the magnetic field
distribution of two magnets, (b) magnetic
forces of different sized NPs, and (c)
NP concentration (20 nm–100 nm) as a
function of distance from the left of the
pillar.

AIP Advances 10, 105209 (2020); doi: 10.1063/5.0023941 10, 105209-2


© Author(s) 2020
AIP Advances ARTICLE scitation.org/journal/adv

FIG. 2. [(a)–(d)] Spatial distributions and


concentration distributions for 20 nm and
40 nm, 20 nm and 60 nm, 20 nm and
80 nm, and 20 nm and 100 nm hybrid
NPs, respectively.

AIP Advances 10, 105209 (2020); doi: 10.1063/5.0023941 10, 105209-3


© Author(s) 2020
AIP Advances ARTICLE scitation.org/journal/adv

⟨Ff (t)⟩ = 0, (5) We simulated the motion of 30 000 magnetic NPs in a pillar
sample and divided the pillar into 40 layers, and the magnetic forces
dr 2 on each magnetic nanoparticle within the polymer matrix were
⟨m( ) ⟩ = 3kB T, (6) obtained at every simulation step. For the micropillars considered,
dt
we adopted the elastic scattering boundary conditions.
where kB is the Boltzmann constant and T is the absolute tempera- As reported earlier,29,36 core–shell NPs (Fe3 O4 @SiO2 , ∼20 nm
ture. in diameter) can be dispersed in the polymer matrices under a

FIG. 3. [(a)–(c)] Spatial distributions and concentration distributions for 100 nm and 40 nm, 100 nm and 60 nm, and 100 nm and 80 nm hybrid NPs, respectively.

AIP Advances 10, 105209 (2020); doi: 10.1063/5.0023941 10, 105209-4


© Author(s) 2020
AIP Advances ARTICLE scitation.org/journal/adv

magnetic field. However, these small particles have two disadvan- As shown in Fig. 2(a), we calculated the spatial distribution and the
tages. First, it requires strong magnetic field excitation and long concentration distribution of 40 nm and 20 nm NPs. According to
driving time to form the gradient due to the relatively low mag- Fig. 1(c), the concentration of the right side is larger than that of the
netic responsivity of small NPs. Second, the ultrahigh specific surface left side for all different sized NPs. However, the right side concen-
areas limit their maximum packing fraction inside the matrix and tration of 20 nm NPs is lower than that of the left side, as shown in
the magnitude of the mechanical gradient of the FGNCs.29 We have Fig. 2(a), in the hybrid NPs. Since 40 nm NPs occupy the right side
made improvements to obtain a concentration gradient distribution space under the large magnetic force, the concentration of 20 nm
of NPs. First, we used two magnets instead of a single magnet to sig- NPs in the right side, therefore, decreases. Different from the pre-
nificantly increase the magnetic force.37 Second, a combination of vious drift model, our model can simulate the space occupied by
hybrid NPs with different ratios was used to maximize the packing NPs. Figures 2(b)–2(d) show the spatial distribution and the con-
fraction inside the polymer matrix. All the parameters were adopted centration distribution of (20 nm and 60 nm), (20 nm and 80 nm),
from our previous work of Refs. 36 and 38. and (20 nm and 100 nm), respectively. Our model did not take into
account the interaction of the particles. The large and small sized
particles move to the right side with the same average velocity under
III. RESULTS the driving of the magnetic field. Since the mass of the large sized
In the model, the distance between the two magnets is 1 cm, and particle is larger than that of the small sized particle, the large sized
the distance from the pillar to the right magnet is 0.2 cm. In addition, particle will have a larger inertial force. With the increase in the NP
the pillar length is 40 μm, and the pillar diameter is 5 μm. Five differ- radius, the concentration of larger NPs also increases. Therefore, the
ent sized NPs (20 nm–100 nm) are selected, as shown in Fig. 1(a). We large size NP will extrude the small size NP at the right side of the
calculated the distribution of the magnetic field, which is shown at sample.
the right of Fig. 1(a). The magnitudes of the magnetic force for NPs As shown in Fig. 3(a), we calculated the spatial distribution and
are shown in Fig. 1(b). Since the pillar size is small compared with the concentration distribution of 100 nm and 40 nm NPs. Since
the distance between two magnets, the magnetic force over the NPs 100 nm NPs occupy the right side space under the large magnetic
was approximated to be uniform. Figure 1(b) shows that the mag- force, the concentration of 40 nm NPs in the right side, therefore,
netic force increases with the particle radius. The large sized NP is decreases. Figures 3(b) and 3(c) show the spatial distribution and
driven by a large magnetic force, while the small sized NP is driven the concentration distribution of (100 nm and 60 nm) and (100 nm
by a small magnetic force at the same position between two mag- and 80 nm), respectively. The large sized NPs extrude the small sized
nets. We choose the concentration of NPs with 40 vol. % inside the NPs at the right side of the sample. It shows that 100 nm NPs squeeze
pillar. Under the driving of the magnetic force, the spatial concen- out the smaller size NPs at the right of the sample. In addition, with
tration distribution was calculated for different sized magnetic NPs, the increase in the NP radius from 40 nm to 80 nm in the hybrid
as shown in Fig. 1(c). The NPs move to the right of the pillar under particles, the concentration of 80 nm is higher than that of 40 nm on
the magnetic field. The concentration distribution of a large sized the right side.
NP (100 nm) has a high gradient under the same magnitude of the To optimize the concentration of NPs, we also investigate 20%
magnetic field. 100 nm NPs mixing with 5%, 10%, 15%, and 20% 20 nm NPs. As
To deeply investigate the concentration distribution of NPs shown in Fig. 4(a), with the increase in the concentration of 20 nm
under the magnetic field, we simulated 20 nm NPs and hybrid dif- NPs, the concentration on the left of the sample increased accord-
ferent sized NPs with an initial concentration of 20% for each size. ingly. In contrast, we simulate 20% 20 nm NPs mixing with 5%,

FIG. 4. (a) NP concentration as a function of distance from the sample left for 100 nm (20 vol. %) and 20 nm (5 vol. %, 10 vol. %, 15 vol. %, and 20 vol. %) and (b) NP
concentration as a function of distance from the sample left for 20 nm (20 vol. %) and 100 nm (5 vol. %, 10 vol. %, 15 vol. %, and 20 vol. %).

AIP Advances 10, 105209 (2020); doi: 10.1063/5.0023941 10, 105209-5


© Author(s) 2020
AIP Advances ARTICLE scitation.org/journal/adv

5
10%, 15%, and 20% 100 nm NPs, as shown in Fig. 4(b). With the B. E. Abali, C. Völlmecke, B. Woodward, M. Kashtalyan, I. Guz, and
increase in the concentration of 100 nm NPs, the concentration on W. H. Müller, Continuum Mech. Thermodyn. 24(4), 377–390 (2012).
6
the right of the sample increased accordingly. Therefore, the con- P. J. Hogg, Science 314(5802), 1100–1101 (2006).
7
centration on the left of the sample can be controlled by tuning the E. D. Wachsman and K. T. Lee, Science 334(6058), 935–939 (2011).
8
initial concentration of 20 nm NPs, and the concentration on the M. Riabkina-Fishman, E. Rabkin, P. Levin, N. Frage, M. P. Dariel, A. Weisheit,
R. Galun, and B. L. Mordike, Mater. Sci. Eng., A 302(1), 106–114 (2001).
right side of the sample can be controlled by tuning the initial con- 9
G. Zuccarello, D. Scribner, R. Sands, and L. J. Buckley, Adv. Mater. 14(18), 1261–
centration of 100 nm NPs. This simulation prediction guides the 1264 (2002).
preparation of various samples and the concentration distribution 10
U. Schulz, M. Peters, F. W. Bach, and G. Tegeder, Mater. Sci. Eng., A 362(1-2),
on both boundaries of the hybrid samples. 61–80 (2003).
11
V. Birman and L. W. Byrd, Appl. Mech. Rev. 60(5), 195–216 (2007).
12
R. K. Singh, D. R. Gilbert, J. Fitz-Gerald, S. Harkness, and D. G. Lee, Science
IV. CONCLUSIONS
272(5260), 396–398 (1996).
13
In short, the spatial concentration gradient of NPs was observed F. Barthelat, Z. Yin, and M. J. Buehler, Nat. Rev. Mater. 1(4) (2016).
14
with an increased particle size. The hybrid NPs were observed to G. N. Pandian and H. Sugiyama, Bull. Chem. Soc. Jpn. 89(8), 843–868 (2016).
15
effectively increase the packing fraction of particles and reduce the B. D. Plouffe, S. K. Murthy, and L. H. Lewis, Rep. Prog. Phys. 78(1), 016601
limitation of the specific surface area of small particles. The desired (2015).
16
concentration distribution can mainly be tuned by adjusting the ini- A. Seidi, M. Ramalingam, I. Elloumi-Hannachi, S. Ostrovidov, and A.
Khademhosseini, Acta Biomater. 7(4), 1441–1451 (2011).
tial concentration and the size of the hybrid NPs. Therefore, the 17
W. Zeng, L. Shu, Q. Li, S. Chen, F. Wang, and X.-M. Tao, Adv. Mater. 26(31),
simulation results provide insight into selecting the NP size and ini- 5310–5336 (2014).
tial concentration in the experiments and show significant benefits 18
J. A. Rogers, T. Someya, and Y. G. Huang, Science 327(5973), 1603–1607 (2010).
in designing functional gradient nanocomposites. 19
K. Autumn, Y. A. Liang, S. T. Hsieh, W. Zesch, W. P. Chan, T. W. Kenny,
R. Fearing, and R. J. Full, Nature 405(6787), 681–685 (2000).
20
V. Imbeni, J. J. Kruzic, G. W. Marshall, S. J. Marshall, and R. O. Ritchie, Nat.
SUPPLEMENTARY MATERIAL Mater. 4(3), 229–232 (2005).
21
Y. P. Bao, T. L. Wen, A. C. S. Samia, A. Khandhar, and K. M. Krishnan, J. Mater.
See the supplementary material for a video simulating Brown-
Sci. 51(1), 513–553 (2016).
ian motion for two sizes of NPs. 22
R. M. Carvalho, A. P. Manso, S. Geraldeli, F. R. Tay, and D. H. Pashley, Dent.
Mater. 28(1), 72–86 (2012).
23
J. De Munck, K. Van Landuyt, M. Peumans, A. Poitevin, P. Lambrechts,
AUTHORS’ CONTRIBUTIONS M. Braem, and B. Van Meerbeek, J. Dent. Res. 84(2), 118–132 (2005).
24
S. Zhang, D. Sun, Y. Fu, and H. Du, Surf. Coat. Technol. 167(2-3), 113–119
D.L. and X.C. contributed equally to this work.
(2003).
25
H. Mei, C. M. Landis, and R. Huang, Mech. Mater. 43(11), 627–642 (2011).
26
C. S. Lee, H. Lee, and R. M. Westervelt, Appl. Phys. Lett. 79(20), 3308–3310
ACKNOWLEDGMENTS
(2001).
27
This work was sponsored by the National Science Foundation M. Zrínyi, Colloid Polym. Sci. 278(2), 98–103 (2000).
28
of China (Grant No. 51972028). X. Z. Xue and E. P. Furlani, Phys. Chem. Chem. Phys. 16(26), 13306–13317
(2014).
29
Z. Wang, X. Shi, H. Huang, C. Yao, W. Xie, C. Huang, P. Gu, X. Ma, Z. Zhang,
DATA AVAILABILITY and L.-Q. Chen, Mater. Horiz. 4(5), 869–877 (2017).
30
E. P. Furlani and X. Xue, Microfluid. Nanofluid. 13(4), 589–602 (2012).
The data that support the findings of this study are available 31
E. P. Furlani, J. Appl. Phys. 99(2), 024912 (2006).
32
from the corresponding author upon reasonable request. S. A. Khashan and E. P. Furlani, Sep. Purif. Technol. 125(125), 311–318 (2014).
33
M. Takayasu, E. Maxwell, and D. Kelland, IEEE Trans. Magn. 20(5), 1186–1188
(1984).
REFERENCES 34
S. Chandrasekhar, Rev. Mod. Phys. 15(1), 1–89 (1943).
1 35
Z. Liu, Y. Zhu, D. Jiao, Z. Weng, Z. Zhang, and R. O. Ritchie, Acta Biomater. 44, G. Friedman and B. Yellen, Curr. Opin. Colloid Interface Sci. 10(3-4), 158–166
31–40 (2016). (2005).
2 36
D. C. Pender, S. C. Thompson, N. P. Padture, A. E. Giannakopoulos, and X. Shi, H. Huang, Z. Wang, and X. Ma, Appl. Sci. 7(11), 1171 (2017).
37
S. Suresh, Acta Mater. 49(16), 3263–3268 (2001). T. Nardi, Y. Leterrier, A. Karimi, and J.-A. E. Månson, RSC Adv. 4(14), 7246
3
S. Suresh, Science 292(5526), 2447–2451 (2001). (2014).
4 38
H. Moshe-Drezner, D. Shilo, A. Dorogoy, and E. Zolotoyabko, Adv. Funct. Z. Wang, K. Wang, D. Liang, L. Yan, K. Ni, H. Huang, B. Li, Z. Guo, J. Wang,
Mater. 20(16), 2723–2728 (2010). X. Ma, X. Tang, and L. Q. Chen, Adv. Mater. 32(25), 2001879 (2020).

AIP Advances 10, 105209 (2020); doi: 10.1063/5.0023941 10, 105209-6


© Author(s) 2020

You might also like