You are on page 1of 8

Available online at www.sciencedirect.

com

ScienceDirect
Procedia Manufacturing 00 (2017) 000–000
www.elsevier.com/locate/procedia

45th SME North American Manufacturing Research Conference, NAMRC 45, LA, USA

Experimental study on the porosity of electrochemical nickel


deposits
Abishek B. Kamaraj1, Hirdayesh Shrestha1, Emily Speck1, and
Murali Sundaram1*†
1
Department of Mechanical and Materials Engineering, University of Cincinnati, Cincinnati, OH, 45220, USA

Abstract

Porous metal parts offer unique advantages over traditional parts as they have excellent specific mechanical properties at a lesser
weight. In this paper, the effect of the electrical parameters of deposition such as voltage and pulse duty cycle during pulsed
electrodeposition on the current density and porosity of the manufactured parts was studied. Porosity at the micron scale, with a
pore size between 1 – 10 µm was identified while using a 250 µm anode (tool). It is demonstrated that the porosity during these
depositions occurs due to the kinetics of the electrochemical deposition, nucleation and crystal growth mechanisms. Through the
study, it was found that the pulse duty cycle and voltage influence the porosity of the part. Higher duty cycle and higher voltage
result in lower porosity in deposits, except in the case of 50% duty cycle when the double layer capacitance confines the deposit.

© 2017 The Authors. Published by Elsevier B.V.


Peer-review under responsibility of the Scientific Committee of NAMRI/SME.

Keywords: Electrochemical Deposition; Porosity; Current Density

1. Introduction

Porous metal parts are increasingly being used in many fields including the energy, environment, metallurgy,
chemical, and biomedical industries [1-4]. The porous part not only inherits the intrinsic metal characteristics such
as weldability, plasticity, thermal conductivity, and electric conductivity, but also displays many new properties
such as small specific weight, controlled permeability, large specific surface area, and high energy absorption [3].

* Corresponding author.
* Dr. Murali Sundaram. Tel.: +1-513-556-2791; fax: +1-513-556-3390.
E-mail address: murali.sundaram@uc.edu

2351-9789 © 2017 The Authors. Published by Elsevier B.V.


Peer-review under responsibility of the Scientific Committee of NAMRI/SME.
2 Kamaraj A. B. et al./ Procedia Manufacturing 00 (2017) 000–000

Some of the methods used in the manufacturing of porous metal parts include injection molding, e-beam melting,
and powder sintering [2, 4, 5]. The recent advances in additive manufacturing have made engineering-controlled
porosity in the part possible. Some of the metal additive manufacturing processes involved in manufacturing porous
parts are selective laser melting (SLM), electron beam melting (EBM), laser engineered net shaping (LENS), and
inkjet 3D printing [6-8]. These methods suffer from lower part quality because of thermal stress or are limited by
the powder size, which determines the minimum feature size [9, 10]. The minimum pore size of the parts using
these processes can be engineered only to about 20 µm. Controllability of the porosity of parts is an issue in
Selective Laser Sintering (SLS) [11]. One parameter that porosity depends on in SLS is the alignment of the powder
particles. SLS is dependent on the motions of the powder, which are both unpredictable and hard to control [12]. In
powder metallurgy, porosity is dependent on the pressure of the gas and the metal’s tension and thus has no
replicability, causing each part produced to be slightly different than others [13]. For the sintering and dissolution
process, there is randomness associated with the template placement, causing the porosity to be unpredictable [13].
A recently developed process combining the principles of AM and Localized Electrochemical Deposition
(LECD), Electrochemical Additive Manufacturing (ECAM) is a novel process capable of producing parts from any
conducting material, such as metals, metal alloys, semiconductors, and conducting polymers [14]. Since it is a non-
thermal process, ECAM is capable of producing parts with low residual stresses and is able to overcome some
challenges of traditional AM processes, such as the need for support structures [14, 15]. Material addition in ECAM
is performed by LECD. In this paper, the porosity of the deposits made by LECD under varying process parameters
was studied. As the LECD process adds material atom by atom through an electrochemical reaction, the porosity
study also gives insights into the electrochemical deposition mechanisms. This increases the controllability of the
porosity achievable in this process.

2. Literature Review

The feasibility of the LECD technique was reported in [16]. Most of the following studies focused on
investigating the effects of various process parameters on the characteristic of the deposited structures and the
deposition rate [17]. While there is a minimum voltage required for deposition to occur, very high voltages cause
porous or irregular deposition structures due to the depletion of ions at high currents and bubble formation [18]. The
same study reported that there was no significant effect of the electrolyte concentration on the deposition rate, but
lower concentrations affected the quality of the deposit (porous) [18]. This phenomenon is again explained due to
the reduction of ions from the formation of the depletion layer at lower electrolyte concentrations. The choice of
electrolyte for the electrodeposition process has been derived mostly from the electrolyte used for the electroplating
of the same metal. Organic additives to the electrolyte produce smooth and fine-grained microcrystalline deposits
[18, 19]. This is due to the altering of the reduction (deposition) mechanism with the addition of additives. Insulation
of the micro tool electrode (anode) results in improved localization of the deposition [16]. However, insulation layer
damage due to the formation of bubbles is an issue as it limits the life of the tool. The accumulation of bubbles,
which hinders the deposition by blocking the interelectrode gap region, is another reason for the reduction in quality
of the deposits. Repeatability of localized electrochemical deposition has been a challenge due to vigorous bubble
formation disturbing the interelectrode dynamics as well as a lack of a proper feedback gap control to maintain a
stable interelectrode gap [20]. In one study, the deposition rate increased toward the center of the cathode, which
caused a more conical shape. The parts produced tend to be more porous toward the bottom of the pillar than the top
[21].
Several studies focusing on the quality of the deposit during electrochemical deposition has been reported in the
literature [17-19, 22]. Studies have shown that several mechanisms and process parameters such as current density,
gas bubbles, feeding mechanism, electrolyte flow, and pulse parameters influence the quality of the electrodeposited
part, but these parameters are all related to the current density associated during the deposition. In one
electroforming study considering three different kinds of wave pulse, it was concluded that a triangular waveform
produced the best surface finish as compared to a sine and a square wave pulse at a higher current density. The range
of current densities studied was 5 to 19.2 A/dm2, and the deposits had a smooth surface finish and a smaller grain
size. A high nucleation rate is promoted by the high pulse current density of the triangular waveform, leading to a
Kamaraj et al./ Procedia Manufacturing 00 (2017) 000–000 3

smaller grain size, and thus less porous deposits [23]. Another study concluded that lowering the current density will
reduce the emission of hydrogen bubbles, thus improving the quality of the deposit [18]. Mechanical methods have
also been used to enhance deposit characteristics and resolution. One method included the utilization of a rotating
electrode [24] which created a uniform deposition field to overcome defects in the electrode tip. Another method
utilized ultrasonic vibrations during deposition [25] to enable the regular removal of air bubbles which accumulate
during deposition and often block deposition growth causing pores. The duty cycle of the pulse power affects the
surface finish of the deposited structure due to the mass transport of the ions during the off time, resulting in the
replenishment of the ions [26, 27]. These studies, however, did not consider nanosecond pulses (high frequency)
which enable the localization of the deposits by reducing the stray current, because of the double layer capacitance
[14, 28].
From the above literature review, claims on the effect of process parameters such as current density and duty
cycle on the quality of the deposits during LECD is inferred. Furthermore, none of the studies quantify the porosity
of the produced deposits, as the aim of those studies was to improve deposit quality. In this paper, we aim to
quantify the amount of porosity generated during the ECAM process and study the effect of voltage and duty cycle
on the current density and the porosity of the parts.

3. Experimental Methods

The schematic of the LECD experimental setup is given in Figure 1. Watts bath containing nickel sulfate (240g),
nickel chloride (45g) and boric acid (30g) per liter of distilled water was used as the electrolyte as this has a lower
throwing power than other nickel deposition electrolytes [29]. Boric acid is used a pH buffer for the bath. While
nickel sulfate provides the majority of the cations needed for the deposition, nickel chloride aids in the deposition of
fine-grained nickel [30]. A side-insulated platinum microelectrode (Ø250 µm) and a polished brass metal plate (2
cm X 2 cm) were used as the anode and cathode, respectively. A synthetic electrical insulating resin and enamel was
used as the insulating material. The current was monitored by measuring the voltage across a 100-ohm resistor
connected in series to the electrical circuit using the Arduino controller. The electrode gap was controlled using a
precision 3 axis stage with a stepper motor resolution of 1.5 µm. The exposed side of the insulated tool was limited
to a length of 500 µm. A pulsed power supply was used to provide current pulses with varying levels of voltages and
duty cycles at 100 ns pulse periods. Duty cycle is defined as the percentage of time the pulse power is on (t on)to the
total time period of the pulse (ton +toff). Five trials of the deposits under each experimental condition were conducted
to minimize the effect of experimental variations. The mass of the samples was measured using a precision weigh
balance, before and after the deposition. The scale had a resolution of 0.1 mg and repeatability of 0.2 mg. Feedback
control with current as the feedback was used to monitor and control the process. The stage and the current
monitoring was done using an in-house built controller. The experimental parameters are listed in Table 1.

Table 1: Deposition Process Parameters

Process Parameter Value

Voltage (V) 4, 5
Duty cycle (%) 50, 75, 100
Initial interelectrode gap (µm); 1–2
Total height of part (µm) 1000
Pulse Period (ns) 100
Tool (anode) diameter (µm) 250
Ni concentration in electrolyte (mol/m3) 1600
4 Kamaraj A. B. et al./ Procedia Manufacturing 00 (2017) 000–000

Figure 1: Schematic of the Localized Electrochemical Deposition Process

Porosity is defined in this study as the ratio of the volume of pores to the total volume of the part as given in
Equation 1.

VP
P  100 (1)
VS  V P

where, P = Porosity percentage, VP = Volume of pores and V S = Volume of Solid. The volume of the deposited
solid was calculated based on the mass of the deposit and density of nickel. The volume of the pores was calculated
as the difference between the total volume of the part and the volume of the solid. The volume of the part was
approximated by building a solid CAD model of the part. The CAD model was built using data from the 2D images
as shown in Figure 2. Smoothening of the extracted boundary was performed so that the CAD software can handle
the data. It is important to note that this method is only an approximation of the volume as it is difficult to exactly
estimate the volume from a single 2D image.

Figure 2: Conversion of 2D images of the part into 3D CAD models to estimate the part volume
Kamaraj et al./ Procedia Manufacturing 00 (2017) 000–000 5

4. Results and Discussion

To understand the microstructural causes of porosity during the ECAM process, the growth of nickel crystals
during the initial stages of the deposition was captured using a scanning electron micrograph (SEM) as shown in
Figure 3 (left). The packing of the growing nickel crystals from the nucleation sites seems to be having an impact on
the inherent porosity of the part. To study the pore size distribution in the parts, the SEM images were analyzed
using an automated watershed segmentation algorithm [31] to detect and separate pores. The relative frequency and
average pore sizes were quantified using this method. Figure 4 shows the image used for the pore size distribution
analysis and the segmented view of the pores. The average pore diameter was calculated to be 6.6 µm. The pore size
distribution of the part is given in Figure 3 (right).

Figure 3: Porosity analysis. (Left) SEM micrograph of deposit shown after 20 seconds of deposition under 5 V and 250-micron tool at 100%
Duty Cycle. The part seen has micron scale porosity showing the growth of the deposits vertically from the nucleation sites. Scale bar size 5 µm.
(Right) Pore size distribution histogram for the deposit shown below.

Figure 4: Top section of porous part (5V 75% D/C) at 500X magnification used for pore size distribution image analysis. (Left) Scanned SEM
image (Middle) Isolated binary image of part (Right) segmented view of the pores found by the algorithm

Some of the deposits under different process conditions are shown in Figure 5. The porosity of these parts was
measured using the method described in the previous section.
6 Kamaraj A. B. et al./ Procedia Manufacturing 00 (2017) 000–000

Figure 5: SEM micrograph of the deposits under various process conditions. Tilt angle 45º and 80X magnification. Scale bar size 200 µm

The effect of duty cycle and the voltage on the current density and porosity of the parts is shown in Figure 6.
Current density values used in this process are noticed to be an order of magnitude higher than the typical
electrochemical studies done for macro scale depositions. Higher voltage and duty cycle result in higher current
density values. The current density values are tending towards limiting current range as seen in the lack of change in
the current density between 75 % and 100 % duty cycle in the 4 V range. The formation of this limiting current
might be due to the depletion of ions at higher overpotentials.

Figure 6: Effect of Voltage and Pulse Duty Cycle (Left) on the Current density; (Right) on porosity during ECAM

The porosity of the parts varied from 20 – 70 %. A duty cycle of 75 % seems to be producing the most porous
parts at both 4 V and 5 V. Compared to the DC potential the pulsed power produces more porous parts. Also, higher
voltages show a reduction in porosity. This shows a trend of lower porosity at higher current densities except for the
case of 50 % duty cycle. At 50 % duty cycle the deposit is smaller in size compared to the tool as seen in the
comparison shown in scanning electron microscope (SEM) images in Figure 5. This might be due to the localization
effect of nanosecond pulses as noted in the literature review section. The stray currents are minimized at smaller
interelectrode gaps leading to confined deposits. This localization is enhanced with only 50 ns of on-time for the
pulses. This phenomenon is seen in both 4 V and 5 V, thus showing a reduction in the porosity of the part using 50
% duty cycles. One reason for the lower porosity at higher current density might be the faster growth of the
individual crystals and increase in nucleation sites at high current densities. At 75 % duty cycle, the deposit growth
is stopped for a small period during the off time and new nucleation starts at a different spot resulting in more porous
parts. This phenomenon is not seen in the 50% duty cycle as the deposit growth is confined to a very small region.
This is due to the elimination of the stray currents by the electrical double layer capacitance.
Kamaraj et al./ Procedia Manufacturing 00 (2017) 000–000 7

5. Conclusion

The porosity of the parts produced during LECD was found to be between 20 – 70 %. Higher current density
realized through higher voltage and higher duty cycle resulted in the least porous parts. Parts manufactured at 75%
duty cycle at 4 V applied potential were the most porous (~ 60 %). The pulsed current forces the deposition to start
at different nucleation sites, resulting in porous parts when pulses are used for the deposition. The pore size
distribution study of the parts shows the scale of the pores formed during this process to be less than 10 µm.

Acknowledgements

Financial support provided by the National Science Foundation under Grant Nos. CMMI-1400800 and CMMI-
1454181 are acknowledged.

References

[1] J.J. Hwang, P.Y. Chen, Heat/mass transfer in porous electrodes of fuel cells, International Journal of Heat and Mass Transfer, 49 (2006) 2315-
2327.
[2] P. Imgrund, A. Rota, F. Petzoldt, A. Simchi, Manufacturing of multi-functional micro parts by two-component metal injection moulding, Int J
Adv Manuf Technol, 33 (2007) 176-186.
[3] P.S. Liu, K.M. Liang, Review Functional materials of porous metals made by P/M, electroplating and some other techniques, Journal of
Materials Science, 36 (2001) 5059-5072.
[4] G. Ryan, A. Pandit, D.P. Apatsidis, Fabrication methods of porous metals for use in orthopaedic applications, Biomaterials, 27 (2006) 2651-
2670.
[5] X. Li, C. Wang, W. Zhang, Y. Li, Fabrication and characterization of porous Ti6Al4V parts for biomedical applications using electron beam
melting process, Materials Letters, 63 (2009) 403-405.
[6] T. Laoui, E. Santos, K. Osakada, M. Shiomi, M. Morita, S. Shaik, N. Tolochko, F. Abe, Properties of titanium implant models made by laser
processing, in: Proc Inst Mech Eng CJ Mech Eng Sci, 2004, pp. 857-863.
[7] O.A. Abdelaal, S.M. Darwish, Fabrication of tissue engineering scaffolds using rapid prototyping techniques, World Academy of Science,
Engineering and Technology, International Journal of Mechanical, Aerospace, Industrial, Mechatronic and Manufacturing Engineering, 5
(2011) 2317-2325.
[8] A. El-Hajje, E.C. Kolos, J.K. Wang, S. Maleksaeedi, Z. He, F.E. Wiria, C. Choong, A.J. Ruys, Physical and mechanical characterisation of
3D-printed porous titanium for biomedical applications, Journal of Materials Science: Materials in Medicine, 25 (2014) 2471-2480.
[9] J.P. Kruth, L. Froyen, J. Van Vaerenbergh, P. Mercelis, M. Rombouts, B. Lauwers, Selective laser melting of iron-based powder, in, Elsevier
Ltd, 2004, pp. 616-622.
[10] M.P. Mughal, H. Fawad, R. Mufti, Finite element prediction of thermal stresses and deformations in layered manufacturing of metallic parts,
Acta Mechanica, 183 (2006) 61-79.
[11] A. Garg, J.S.L. Lam, M.M. Savalani, A new computational intelligence approach in formulation of functional relationship of open porosity
of the additive manufacturing process, The International Journal of Advanced Manufacturing Technology, 80 (2015) 555-565.
[12] E.J.R. Parteli, T. Pöschel, Particle-based simulation of powder application in additive manufacturing, Powder Technology, 288 (2016) 96-
102.
[13] M. Covaciu, M. Walczak, J. Ramos-Grez, A method for manufacturing cellular metals with open- and close-type porosities, Materials
Letters, 65 (2011) 2947-2950.
[14] M.M. Sundaram, A.B. Kamaraj, V.S. Kumar, Mask-less Electrochemical Additive Manufacturing: A Feasibility Study, Journal of
Manufacturing Science and Engineering, 137 (2015) 021006-021015.
[15] A. Brant, M. Sundaram, A Novel Electrochemical Micro Additive Manufacturing Method of Overhanging Metal Parts without Reliance on
Support Structures, Procedia Manufacturing, 5 (2016) 928-943.
[16] J.D. Madden, I.W. Hunter, Three-dimensional microfabrication by localized electrochemical deposition, Microelectromechanical Systems,
Journal of, 5 (1996) 24-32.
[17] R.A. Said, Localized electro-deposition (LED): The march toward process development, Nanotechnology, 15 (2004) S649-S659.
[18] E.M. El‐Giar, R.A. Said, G.E. Bridges, D.J. Thomson, Localized Electrochemical Deposition of Copper Microstructures, Journal of The
Electrochemical Society, 147 (2000) 586-591.
[19] A. Jansson, G. Thornell, S. Johansson, High Resolution 3D Microstructures Made by Localized Electrodeposition of Nickel, Journal of The
Electrochemical Society, 147 (2000) 1810-1817.
[20] R.A. Said, Adaptive tip-withdrawal control for reliable microfabrication by localized electrodeposition, Journal of Microelectromechanical
Systems, 13 (2004) 822-832.
[21] H. Xiao, P. Zeng, X. Ren, F. Wang, Three-dimensional microfabrication of copper column by localized electrochemical deposition, in: 2016
17th International Conference on Electronic Packaging Technology (ICEPT), 2016, pp. 69-72.
8 Kamaraj A. B. et al./ Procedia Manufacturing 00 (2017) 000–000

[22] T.K. Chang, J.C. Lin, J.H. Yang, P.C. Yeh, D.L. Lee, S.B. Jiang, Surface and transverse morphology of micrometer nickel columns
fabricated by localized electrochemical deposition, Journal of Micromechanics and Microengineering, 17 (2007) 2336-2343.
[23] K.C. Chan, W.K. Chan, N.S. Qu, Effect of current waveform on the deposit quality of electroformed nickels, Journal of Materials Processing
Technology, 89–90 (1999) 447-450.
[24] S.H. Yeo, J.H. Choo, Effects of rotor electrode in the fabrication of high aspect ratio microstructures by localized electrochemical
deposition, Journal of Micromechanics and Microengineering, 11 (2001) 435-442.
[25] S.H. Yeo, J.H. Choo, K.H.A. Sim, On the effects of ultrasonic vibrations on localized electrochemical deposition, Journal of
Micromechanics and Microengineering, 12 (2002) 271-279.
[26] J.C. Lin, T.K. Chang, J.H. Yang, Y.S. Chen, C.L. Chuang, Localized electrochemical deposition of micrometer copper columns by pulse
plating, Electrochimica Acta, 55 (2010) 1888-1894.
[27] M.A. Habib, S.W. Gan, M. Rahman, Fabrication of complex shape electrodes by localized electrochemical deposition, Journal of Materials
Processing Technology, 209 (2009) 4453-4458.
[28] R. Schuster, V. Kirchner, P. Allongue, G. Ertl, Electrochemical Micromachining, Science, 289 (2000) 98-101.
[29] Z. Abdel-Hamid, Improving the throwing power of nickel electroplating baths, Materials Chemistry and Physics, 53 (1998) 235-238.
[30] W.A. Wesley, J.W. Carey, The Electrodeposition of Nickel from Nickel Chloride Solutions, Transactions of The Electrochemical Society,
75 (1939) 209-236.
[31] A. Rabbani, S. Jamshidi, S. Salehi, An automated simple algorithm for realistic pore network extraction from micro-tomography images,
Journal of Petroleum Science and Engineering, 123 (2014) 164-171.

You might also like