You are on page 1of 6

International Journal of Modern Physics B

Vol. 33 (2020) 2040038 (6 pages)


c World Scientific Publishing Company
DOI: 10.1142/S021797922040038X
by UNIVERSITY OF AUCKLAND on 01/10/20. Re-use and distribution is strictly not permitted, except for Open Access articles.

Cu–TiO2 nanocomposite coatings prepared


from sol-enhanced electrodeposition

Zhen He∗,§,k , Shengping Zhang∗ , Li Yin∗ ,


Muhammed D. Hayat† and Peng Cao†,‡,¶,k
∗ School
Int. J. Mod. Phys. B Downloaded from www.worldscientific.com

of Materials Science and Engineering,


Jiangsu University of Science and Technology,
Zhenjiang 212003, P. R. China
† Department of Chemical and Materials Engineering,

University of Auckland, Private Bag 92019, Auckland 1142, New Zealand


‡ MacDiarmid Institute for Advanced Materials and Nanotechnology,

University of Victoria Wellington, P. O. Box 600, Wellington, New Zealand


§ hezhen0826@outlook.com
¶ p.cao@auckland.ac.nz

Received 30 April 2019


Accepted 20 August 2019
Published 19 December 2019

A series of TiO2 -modified Cu coatings were developed via a sol-enhanced electrodeposi-


tion process. The effect of TiO2 sol on the properties of fabricated Cu–TiO2 nanocom-
posites is studied. Phase composition and microstructures of Cu–TiO2 deposits were
characterized using X-ray diffraction (XRD), atomic force microscopy (AFM) and trans-
mission electron microscopy (TEM). Mechanical properties of the fabricated Cu–TiO2
coatings were also examined. The proper addition of TiO2 sol gives rise to refined coating
surface with well-distributed TiO2 nanoparticles, leading to significantly enhanced prop-
erties. The microhardness and wear resistance of the Cu–TiO2 nanocomposite coating
(12 mL/L TiO2 sol) upsurges owing to the strengthening effect of highly dispersed TiO2
nanoparticles. We also suggest that the excessive addition of TiO2 sol should be avoided
as it will result in an inferior coating quality with a loose structure for the prepared
Cu–TiO2 deposits.

Keywords: Nanocomposite; TiO2 sol; electrodeposition; copper.

PACS numbers: 68.35.Gy

1. Introduction
Along with the growing industrial demand for high-performance coating materials,
current research is focused on the enhancement of surface properties of applied
engineering coatings.1–5 Owing to its unique combination of low cost and good

k Corresponding authors.

2040038-1
Z. He et al.

electrical- and thermal-conductivity, copper is an extensively applied metallic coat-


ing material in a broad range of industrial fields such as electronic engineering and
thermal conduction engineering.6,7 Nevertheless, the scale-up application of copper
coatings is limited by its poor hardness, low wear resistance and inferior corrosion
by UNIVERSITY OF AUCKLAND on 01/10/20. Re-use and distribution is strictly not permitted, except for Open Access articles.

resistance under harsh service conditions. Recent research has proposed different
methods to modify copper coatings, such as thermomechanical treatment, alloying
and composite formation. Among these, electrodeposition is a promising technique
to control the properties of copper layers by incorporating a suitable second phase
material. Some reports found that incorporating SiC particles gives rise to en-
hanced microhardness in Cu–SiC composite coatings.8,9 Thiemig and co-workers
also confirmed modified surface morphology and microhardness of the electrode-
posited Cu–Al2 O3 composite coatings.10
Int. J. Mod. Phys. B Downloaded from www.worldscientific.com

Many studies have attempted to embed TiO2 nanoparticles into electrodeposited


composite coatings, achieving encouraging improvements in coating microhard-
ness, wear resistance and corrosion resistance, etc.2,11,12 More recently, a novel
sol-enhanced electrodeposition technique was developed to manufacture composite
coatings involving pre-synthesized TiO2 sol, which leads to in-situ formation of TiO2
nanoparticles throughout the co-electrodeposition process.13,14 This sol-enhanced
electrodeposition can improve the dispersion of embedded TiO2 nanoparticles in
the deposited composite coatings. For instance, the use of TiO2 sol in Ni–P and
Au–Ni electrodeposition has been proven to result in impressive enhancement in
the mechanical performance of nanocomposite coatings.13,15 However, no study
has ever exploited the potential of the sol-enhanced electrodeposition technique in
preparing Cu–TiO2 coatings. Herein, we systematically studied the electrodeposited
Cu–TiO2 nanocomposite coatings with the usage of TiO2 sol, aiming to establish
a new method for making Cu–TiO2 coatings with superior performances.

2. Experimental
Analytical grade chemicals and deionized (DI) water were used in sample prepa-
ration. Stainless-steel substrates were prepared by several pretreatment steps: (1)
grinding and polishing; (2) ultrasonic cleaning in acetone and (3) chemical etching
by immersing in the concentrated hydrochloric acid solution. TiO2 sol was prepared
through the methods reported in our previous work.13
The electrodeposition of Cu–TiO2 nanocomposite coatings was conducted in a
beaker cell at a constant current of 40 mA/cm2 at room temperature. The stainless-
steel substrate (20 mm × 30 mm) serves as the anode, while a pure copper sheet
(40 mm × 40 mm) was the cathode. The inter-electrode distance was ∼40 mm,
and the bath agitation was kept at 300 pm. The electrolyte was made by 0.5 M
CuSO4 +0.1 M H2 SO4 . Before the electrodeposition, different amounts of prepared
TiO2 sol were carefully added into the electrolytic bath under magnetic stirring.
The crystal structure of Cu–TiO2 coatings was identified by X-ray Diffraction
(XRD, D2 Phaser, Bruker, Germany) at a step size of 0.02◦ . The surface morphology

2040038-2
Cu–TiO2 nanocomposite coatings prepared from sol-enhanced electrodeposition

of coatings was determined using laser scanning confocal microscopy (Axioskop2-


MAT, Zeiss, Germany) and atomic force microscopy (AFM, Next II, NT-MDT,
Russia). Microstructures were identified by a high-resolution transmission electron
microscopy (TEM, TECNAI G2 F20, FEI, USA).
by UNIVERSITY OF AUCKLAND on 01/10/20. Re-use and distribution is strictly not permitted, except for Open Access articles.

The hardness measurement was performed using a Vickers microhardness test


machine. Five indentations were randomly conducted on the surface of each sample,
and a load of 50 g was applied for 15 s for each indentation. The tribological
tests were carried out employing a micro-tribometer (UMT-2, CETR, USA) at
room temperature. In each measurement, the wear test was conducted applying a
load of 7 N and total sliding distance of 54 m.

3. Results and Discussion


Int. J. Mod. Phys. B Downloaded from www.worldscientific.com

Figure 1 presents the crystal structure of electrodeposited Cu–TiO2 coatings. All


diffraction peaks can be assigned to copper and a significant texture appears at
∼40◦ . There is no TiO2 diffraction peak identified in Fig. 1, because the content of
TiO2 generated in the co-deposition process is too low. Meanwhile, the peak inten-
sity and width change with the amount of TiO2 sol addition. An increased TiO2
sol leads to broader peaks with less intensity until the TiO2 sol reaches 12 mL/L,
beyond which an opposite tendency is observed with increasing TiO2 sol addition.
Figure 2 presents the surface morphology of the Cu–TiO2 nanocomposite coat-
ings. No crack can be seen in these deposits, indicating a good electrodeposition
quality. From Figs. 2(a) to 2(d), an increased TiO2 sol addition significantly refines

Fig. 1. (Color online) XRD patterns of Cu–TiO2 nanocomposite coatings prepared at different
TiO2 sol addition.

2040038-3
Z. He et al.
by UNIVERSITY OF AUCKLAND on 01/10/20. Re-use and distribution is strictly not permitted, except for Open Access articles.
Int. J. Mod. Phys. B Downloaded from www.worldscientific.com

Fig. 2. Surface morphologies of Cu–TiO2 coatings prepared at different addition of TiO2 sol:
(a) 0 mL/L, (b) 4 mL/L, (c) 8 mL/L, (d) 12 mL/L, (e) 16 mL/L and (f) 20 mL/L.

Fig. 3. (Color online) AFM profiles of Cu–TiO2 nanocomposite coatings prepared at different
TiO2 sol addition, (a) 0 mL/L, (b) 12 mL/L and (c) 20 mL/L.

the coating surface, revealing a rather fine and compact surface morphology at
12 mL/L TiO2 sol addition [Fig. 2(d)]. However, an excessive TiO2 sol addition at
20 mL/L causes an inferior coating quality with readily identified crystal clusters,
Fig. 2(f).
Surface quality of the coating is further revealed by AFM and shown in Fig. 3.
When the TiO2 sol increased to 12 mL/L, the surface became smoother [Fig. 3(b)].
However, further increasing TiO2 sol caused a rougher and more open surface mor-
phology, Fig. 3(c). In the electrodeposition process of Cu–TiO2 coatings, the well-
dispersed TiO2 nanoparticles facilitate the nucleation process, thus leading to a
refined surface morphology with decreased crystal size. On the other hand, the ex-
cessive addition of TiO2 sol may result in aggregation of nanosized particles, which
is associated with the inferior coating quality with a rough and open surface.
Neither the structural or morphologic characterization confirm the presence of
nanosized TiO2 due to the detection limit of XRD or SEM, which agrees well with

2040038-4
Cu–TiO2 nanocomposite coatings prepared from sol-enhanced electrodeposition
by UNIVERSITY OF AUCKLAND on 01/10/20. Re-use and distribution is strictly not permitted, except for Open Access articles.
Int. J. Mod. Phys. B Downloaded from www.worldscientific.com

Fig. 4. Microstructure of Cu–TiO2 nanocomposite observed by TEM.

another study.15 The nanosized particles in the composite coating are difficult to
be identified by conventional characterization methods such as XRD and SEM.
Here we used a high-resolution TEM to address this challenge and the result is
shown in Fig. 4. It can be clearly seen that TiO2 nanoparticles are distributed in
the composite coating with a size of several nanometers. The particle size is much
smaller compared with other literature studying nanoparticle-doped composite.16
This is associated with the continuous in situ formation of TiO2 nanoparticles with
a rather fine size, during the electrodeposition process involving pre-synthesized
TiO2 sol.
The hardness and wear resistance were measured for the prepared Cu–TiO2
nanocomposite coatings, which is summarized in Fig. 5. The coating hardness in-
creases from 106 HV (pure copper layer) to 145 HV upon TiO2 sol addition of
12 mL/L. Further increasing the TiO2 sol addition harms the coating hardness.

Fig. 5. (Color online) Microhardness and friction coefficient of Cu–TiO2 nanocomposite coatings
prepared at different TiO2 sol addition.

2040038-5
Z. He et al.

Also, wear resistance measurement reveals a similar trend. The friction coefficient of
Cu–TiO2 coatings gradually drops from 0.84 of pure copper to 0.19 of Cu-12 mL/L
TiO2 sol, and then increases to 0.46 of Cu-20 mL/L TiO2 sol. The best mechanical
behavior is confirmed for the Cu-12 mL/L TiO2 sol nanocomposite coating.
by UNIVERSITY OF AUCKLAND on 01/10/20. Re-use and distribution is strictly not permitted, except for Open Access articles.

At a proper content of TiO2 sol, the in situ generated TiO2 nanoparticles are
highly dispersed in the coating matrix as a second phase. The uniformly distributed
TiO2 nanoparticles in the coating lead to dispersion strengthening effect and grain-
refinement strengthening effect, therefore resulting in the improved hardness and
wear resistance. However, the increasing agglomeration of TiO2 particles in the
coating causes decreased strengthening effect and a loose structure, which negatively
affects mechanical performance.
Int. J. Mod. Phys. B Downloaded from www.worldscientific.com

4. Conclusions
This study pioneered TiO2 sol addition method as an effective approach to prepare
Cu–TiO2 nanocomposite coatings. A greater content of TiO2 sol upto 12 mL/L
resulted in a more refined coating surface with a smaller crystal size. The trend
reversed beyond 12 mL/L and a 20 mL/L addition led to a nonuniform surface
morphology. The 12 mL/L concentration of TiO2 sol was recommended for optimal
mechanical properties of the coatings.

References
1. P. Y. Guo et al., Coatings 8, 11 (2018).
2. W. Gao et al., Surf. Coat. Technol. 350, 801 (2018).
3. Z. He et al., J. Electrochem. Soc. D 165, 670 (2018).
4. Y. Wang et al., Surf. Coat. Technol. 349, 217 (2018).
5. W. H. Zhang et al., Int. J. Mod. Phys. B 33, 8 (2019).
6. M. Hasegawa et al., Electrochim. Acta 178, 458 (2015).
7. P. T. Lee et al., Surf. Coat. Technol. 320, 559 (2017).
8. M. Lekka et al., Surf. Coat. Technol. 205, 3438 (2011).
9. A. K. Pradhan and S. Das, Metall. Mater. Trans. A 45, 5708 (2014).
10. T. Denny, L. Ronny and B. Andreas, Electrochim. Acta 52, 7362 (2007).
11. Y. Wang et al., J. Alloy. Compd. 649, 222 (2015).
12. M. H. Allahyarzadeh et al., J. Alloy. Compd. 705, 788 (2017).
13. Y. Wang et al., J. Electrochem. Soc. D 161, 775 (2014).
14. Y. Wang et al., J. Alloy. Compd. 617, 472 (2014).
15. Y. Wang et al., Int. J. Electrochem. Sci. 9, 4384 (2014).
16. J. Kong et al., Electrochim. Acta 53, 2048 (2007).

2040038-6

You might also like