You are on page 1of 13

Journal of Manufacturing Processes 41 (2019) 10–22

Contents lists available at ScienceDirect

Journal of Manufacturing Processes


journal homepage: www.elsevier.com/locate/manpro

Effect of tool-sidewall outlet hole design on machining performance in T


electrochemical mill-grinding of Inconel 718

Shen Niua, Ningsong Qua,b, , Xiaokang Yuea, Hansong Lia,b
a
College of Mechanical and Electrical Engineering, Nanjing University of Aeronautics and Astronautics, Nanjing, 210016, China
b
Jiangsu Key Laboratory of Precision and Micro-Manufacturing Technology, Nanjing, 210016, China

A R T I C LE I N FO A B S T R A C T

Keywords: Electrochemical mill-grinding (ECMG) combines electrochemical milling and electrochemical grinding in a
Electrochemical mill-grinding single processing procedure, and can be used for both rough and finish machining of difficult-to-machine alloys.
Tool-sidewall outlet holes In the rough machining stage, improved tool designs, obtained by varying the number of rows of the tool-
Inconel 718 sidewall outlet holes and their mode of arrangement, are both introduced for the ECMG of Inconel 718. Four
Material removal rate
tools with different numbers of rows of tool-sidewall outlet holes are designed. The test results show that a
Flatness
Surface roughness
higher maximum feed rate can be obtained by using an abrasive tool with four rows of tool-sidewall outlet holes.
Experimental results on machining a slot with this tool indicate that the material removal rate is increased at
higher applied voltage, electrolyte pressure, and feed rate, while the average slot width and sidewall flatness
become smaller at higher electrolyte pressure and feed rate at fixed applied voltage. However, a major defect of
the vertical alignment of tool-sidewall outlet holes is found to be the formation of an uneven profile on the
machined sidewall, which gives rise to a large machining allowance for subsequent finishing. Therefore, an
abrasive tool with a spiral arrangement of tool-sidewall outlet holes is proposed. The average sidewall flatness
and the sidewall surface roughness obtained with the original tool are 549.6 μm and 2.509 μm, but they are only
340.5 μm and 1.65 μm with the new tool. In addition, the new tool is also applied for finish machining on the slot
sidewall produced by rough machining. After the finish machining stage, the average sidewall flatness and the
sidewall surface roughness decrease from 340.5 μm to 69.5 μm and from 1.65 μm to 0.648 μm, respectively.

1. Introduction ECG, a soft, non-reactive oxide layer can occur on the workpiece surface
by electrolytic reaction, and then can be immediately removed by the
Inconel 718 is a nickel alloy with superior mechanical properties at cutting action of abrasive grains, thus exposing fresh metal for continue
high temperature that make it one of the most commonly used struc- electrolytic reaction [10–12]. In this way, the process productivity is
tural materials for aero-engine components such as blisks and thin- increased many times while tool wear is significantly reduced com-
walled casings [1–3]. However, it is well known that Inconel 718 is pared with conventional grinding, particularly when working with
difficult to cut because of its high shear strength and low thermal difficult-to-cut materials [13–15]. Zaborski et al. [16] found that the
conductivity, which lead to severe tool wear, low productivity, and wear on a diamond grinding wheel in mechanical grinding of sintered
poor surface integrity during conventional machining processes [4–6]. carbides and titanium alloy was about 15 times greater than in ECG.
The presence of many complex structures in aero-engine components Goswami et al. [17] observed that there was about a 75–95% reduction
further increases the difficulty of manufacturing [7,8]. These limita- in grinding force using ECG compared with conventional grinding for
tions mean that conventional machining processes do not meet the machining an Al2O3/Al interpenetrating-phase composite. In addition,
demands of the modern aerospace industry. Hence, it is important to ECG also has the ability to finish a metal material of any strength and
find a method for machining Inconel 718 with low cost and high effi- hardness, giving a burr-free surface and with no thermal or mechanical
ciency and flexibility. damage to the workpiece [18,19]. Roy et al. [20] carried out a detailed
Electrochemical grinding (ECG) is a nonconventional hybrid ma- study on the effect of voltage on surface texture in ECG and found that
chining process that combines the capabilities of electrochemical ma- electrochemical action contributed about 90% of the material removal.
chining (ECM) and mechanical grinding in a single operation [9,10]. In Hasçalık and Çaydaş [21] reported that the ECG process effectively


Corresponding author at: College of Mechanical and Electrical Engineering, Nanjing University of Aeronautics and Astronautics, Nanjing, 210016, China.
E-mail address: nsqu@nuaa.edu.cn (N. Qu).

https://doi.org/10.1016/j.jmapro.2019.03.027
Received 8 August 2018; Received in revised form 27 February 2019; Accepted 22 March 2019
1526-6125/ © 2019 The Society of Manufacturing Engineers. Published by Elsevier Ltd. All rights reserved.
S. Niu, et al. Journal of Manufacturing Processes 41 (2019) 10–22

eliminated damaged surface layers and improved the surface integrity


of Ti–6Al–4 V alloy machined by electrical discharge machining.
The major problem of ECG when machining complex parts is a lack
of manufacturing flexibility. In the studies already cited, ECG typically
involves the use of a large (diameter 150–350 mm) grinding wheel, and
the thickness of material removed in a single pass is shallow (depth of
cut 0.005–0.12 mm). Although few in number, some recent papers have
described in detail the use of smaller-diameter abrasive tools in the
motion mode of milling, for improving the machining flexibility and
efficiency of ECG. Qu et al. [22] applied a ball-ended rod-shaped dia-
mond wheel to ECG of Inconel 718 with X–Y–Z linear motion. They
obtained a material removal rate (MRR) of 85.8 mg min−1
(10.41 mm3 min−1) at a depth of cut of 0.5 mm. Li et al. [23,24] pro-
posed a method of inner-jet ECG of the nickel-based superalloy GH4169
utilizing an abrasive tool based on a dead-end tube from which the
electrolyte was directly ejected through six tool-sidewall outlet holes. In
Fig. 1. Schematic diagram of electrolyte flow in ECMG with a depth of cut of
their studies, the MRR was increased to 0.493 g min−1 3 mm.
(59.83 mm3 min−1) by using a depth of cut of 3 mm, and samples with
convex structure were also produced using a planned milling path. Niu
et al. [25] termed this modified ECG process electrochemical mill- 2. Design of tool-sidewall outlet holes for a tool substrate
grinding (ECMG). They fabricated a sample with a thin-walled
(0.5 mm) structure at a depth of cut of 3 mm using the above inter-jet 2.1. Four tool substrates with different numbers of rows of tool-sidewall
tool with bottom insulation. outlet holes
Unlike the conventional ECG process, ECMG combines electro-
chemical (EC) milling and ECG in a single processing procedure, and According to previous experimental results, a tool with single row of
therefore is able to process components with complicated shapes. tool-sidewall outlet holes was recommended for ECMG, in order to
Furthermore, ECMG can also be used for both rough and finish ma- achieve a high MRR at a depth of cut of 3 mm (see Fig. 1) [23–25].
chining without the need to replace the tool and re-fix the workpiece. In Although a greater depth of cut is beneficial for a higher MRR, it also
the stage of rough machining, ECMG is like EC milling, and its main makes rapid removal of processing by-products from the frontal gap
goal is to achieve a high MRR by enhancing electrochemical dissolution more difficult. The accumulation of processing by-products in the
[26]. In the stage of finish machining, ECMG is like ECG, and its major frontal gap leads to an accelerating deterioration in electrochemical
purpose is to obtain high machining accuracy and surface quality by dissolution, which can cause a significant decline in machining stability
enhancing the grinding action [27,28]. Thus, ECMG with appropriate and product quality. Therefore, when a larger depth of cut is required,
parameter settings can replace two separate machining operations. In the use of a greater number of rows of tool-sidewall outlet holes should
fact, the requirement for both rough and finish machining arises in the be considered, with the aim of providing more powerful flushing in the
manufacture not only of components for aero-engines but also those for frontal gap.
helicopter transmissions, such as rotor shafts, rotor hubs, and con- In this study, the outer diameter of the tool substrate and its desired
nectors. Hence, ECMG has great potential for application in the aircraft depth of cut are both 10 mm. Initial investigations are conducted to
industry. determine the number of rows of tool-sidewall outlet holes. As shown in
Rough machining is the first step in the production of components, Fig. 2, four tool substrates are designed (tools A, B, C, and D), in which
in which a volume of material needs to be removed for the subsequent there are respectively two, three, four, and five rows of outlet holes. In
finish machining. Based on the studies mentioned above, it is clear that all four tool substrates, the distance between the hole center of the first
the MRR of ECMG tends to be related to the depth of cut, which in turn row and the bottom of the tool substrate is 1.6 mm, while the distance
is determined by the use of an inter-jet tool. Hence, a rational design of between the hole centers of any two adjacent rows is 6.8, 3.4, 2.3, and
the tool-sidewall outlet holes is a prerequisite for efficient removal of a 1.7 mm in tools A, B, C, and D, respectively. The outlet hole diameter
large machining allowance in one pass. It is worth recalling that pre- and tube wall thickness of each tool substrate are both 1 mm, the radius
vious studies have only considered the effect of the tool-sidewall outlet of the rounded corner at the tool’s bottom edge is 0.5 mm, and there are
holes on the flow field, while neglecting their harmful impact on the six outlet holes in each row.
electric field [23]. In fact, machining performances such as the MRR
and the flatness and surface quality of the machined sidewall are de- 2.2. Flow field analysis
termined by the integrated effects of these two physical fields. How-
ever, there is still a lack of relevant studies on the effect of the design of Flow field simulation is carried out to clarify the effect of the design
tool-sidewall outlet holes on the above machining performances. Ad- of tool-sidewall outlet holes on the electrolyte flow process. In this
ditionally, there have been no reports of finish machining in ECMG.
In this paper, to enhance the machining performances in ECMG of
Inconel 718, different inter-jet tools are designed by varying the
number of rows of tool-sidewall outlet holes and their mode of ar-
rangement, and the factors influencing the uniformity of the flow field
and that of the electric field by are revealed by numerical simulations
and experimental verifications. Furthermore, a series of comparative
trials are conducted to analyze the effects of different processing
parameters and abrasive tools on the MRR and the flatness and surface
quality of the machined sidewall. Finally, finish machining using the
optimal abrasive tool is also performed to improve the flatness and
surface quality of the machined sidewall produced by rough machining. Fig. 2. Four tool substrates with different numbers of rows of tool-sidewall
outlet holes.

11
S. Niu, et al. Journal of Manufacturing Processes 41 (2019) 10–22

low flow velocity are found in the frontal gap, compared with the case
of tool C. A reasonable explanation is that the quantity of electrolyte
draining away from the right side of the outlet holes increases notice-
ably when the number of rows of tool-sidewall outlet holes exceeds a
certain value. As a result, the flow velocity in the left side of the outlet
holes is reduced, which gives rise to a deterioration in flow field uni-
formity in the frontal gap.

2.3. Electric field analysis

Because the conductive area of the tool sidewall decreases with


increasing number of rows of outlet holes, a static simulation of the
electric field needs to be performed to analyze the effect of the different
tool substrates on the current density distribution at the machining
surface of the slot. A geometric model for use in the numerical simu-
lation is shown in Fig. 5. The electric field is the interspace between the
Fig. 3. Model geometry of the electrolyte flow domain in the numerical simu- tool substrate and the slot [25]. In this geometric model, the inter-
lation. electrode gap is 0.2 mm and the slot length is 7 mm. The following
conditions and assumptions are adopted:
study, the rotation of the tool substrate in the fluid environment is si-
mulated using the sliding mesh technique [29]. The following condi- 1 The electrolyte conductivity remains approximately constant.
tions and assumptions are adopted: 2 The current density distribution on the anode faces is determined
only by Ohmic effects.
1 The electrolyte flow is incompressible and continuous. 3 The surfaces of the cathode tool and the workpiece are taken as
2 There are no bubbles or solid particles in the electrolyte flow. equipotential surfaces.
3 Solid surfaces are no-slip walls.
4 Energy dissipation and electrolyte temperature changes are ignored. The electrical potential φ in the computational domain is described
5 The electrolyte flow satisfies the mass and momentum conservation by the Laplace equation
equations. ∇2 φ = 0 (1)
The boundary conditions are taken as
An example of the geometry of the electrolyte flow field is shown in
Fig. 3. The flow field is the interspace between the tool substrate and φ|Γw = U (workpiece) (2)
the slot [23]. In the geometric model, the inter-electrode gap is 0.2 mm
and the slot length is 7 mm. The geometric model is further divided into φ|Γc = 0 (cathode tool) (3)
a stationary zone and a rotating zone, and the interfaces are defined to
∂φ
interchange the date between two zones. Moreover, two vertical sec- =0 (free boundaries)
∂n (4)
tions are chosen to observe the flow velocity distributions, with section Γf

A being located in the middle of the slot and section B 0.05 mm distant where U is the applied voltage and n the is unit normal vector to the
from the slot sidewall. surface. According to Ohm’s law, the relationship between the current
Unstructured tetrahedral meshes with an element size of 0.1 mm are density i and the electrical potential φ is given by
employed for discretization of the electrolyte flow domain to solve the
conservation equations of mass and momentum. On the basis of the U UκA
I= =
R Δ (5)
actual experimental conditions, the inlet pressure is 0.6 MPa, the outlet
pressure is the pressure of the atmospheric environment (0.1 MPa), and I Uκ
the rotational speed of the rotating zone is 1000 rev min−1. The stan- i= = = κ∇φ
A Δ (6)
dard κ–ε turbulence model is chosen to solve the Reynolds-averaged
where R is the equivalent resistance, κ is the electrolytic conductivity,
Navier–Stokes equation, which is appropriate for the turbulent flow in
and Δ is the interelectrode gap
the 3D model of ECM [23]. The flow field simulation is carried out with
For all cases, COMSOL 5.2 simulation software was used to calculate
a time step of 0.00025 s using the computational fluid dynamics soft-
the results for the corresponding experimental conditions, with U =
ware FLUENT 15.
30 V and κ = 10 S m−1.
Fig. 4 shows the velocity contour at section A, which illustrates the
Fig. 6 shows the current density distribution over the machining
electrolyte flow situation at the frontal gap. A large difference in flow
surface of the workpiece for different tool substrates. It can be seen
velocity can be seen clearly between the outlet holes on the left and
from the simulation results that the current density provided by the
right sides due to the high fluid resistance of the frontal gap, which
cathode surface reaches as high as 150 A cm−2, but the current density
results in insufficient flushing in this gap. When tool A is used (Fig. 4a),
in the region opposite the tool-sidewall outlet hole is only about 70 A
a long dead-water zone occurs at the middle of the frontal gap. With
cm−2. This is because there is a lack of cathode surface in the neigh-
tool B (Fig. 4b), there is still a dead-water region at the frontal gap
borhood of the outlet hole, which gives rise to a low electric field in-
between any two adjacent outlet holes, but its area is significantly less
tensity on the region opposite the outlet hole. According to Eq. (6), the
than in the case of tool A. From Fig. 4c, the employ of tool C can further
current density depends on the electric field intensity. As a result, a
decrease the area of each dead-water region in the frontal gap in con-
circular region with low current density occurs on the machining sur-
trast to the case of tool B. This is mainly because, with an increase in the
face in front of the outlet hole. Furthermore, from Faraday’s law, the
number of rows of tool-sidewall outlet holes, the flushing ability in the
metal dissolution rate will be proportional to the current density. From
frontal gap is effectively enhanced, which leads to an improvement in
Fig. 6, it can be also seen that the number of regions of low current
flow field uniformity. However, when tool D is applied (Fig. 4d), not
density in the vertical direction increases with the number of rows of
only does the dead-water region still exist but also more regions with
outlet holes, which also means a greater numbers of regions where the

12
S. Niu, et al. Journal of Manufacturing Processes 41 (2019) 10–22

Fig. 4. Flow velocity distributions with different tool substrates at section A.

controlled by a computerized numerical control system consisting of a


servo motion controller and four-axis motion components. The rotary
joint and the slip ring were connected to the rotating spindle, allowing
internal flushing and current conduction, respectively. The machining
current was continuously monitored and recorded by a data acquisition
card interfaced with a computer. The hydrogen generated by electro-
chemical dissolution was discharged promptly via ventilation. In ad-
dition, an electrolyte circulation and filtration unit was designed for
supplying clear electrolyte with the desired pressure and temperature.
A power supply was used to provide the applied voltage required for
ECMG.
An electroplating process was employed to fabricate the
Ni–diamond tools, in which the diamond particle size and concentra-
tion were 75–90 μm and 8.8 carat cm−3, respectively. The four types of
Fig. 5. Model geometry of the electric field in the numerical simulation.
abrasive tools, each of external diameter about 10.2 mm, were shown in
Fig. 8. The density of workpiece specimen made of Inconel 718 was
8.24 g cm−3. In this paper, the set depth of cut was 10 mm in a single
dissolution rate is slower. This may threaten machining stability. pass. In addition, the electrolyte with a temperature of 30 °C was a 10%
Hence, it is necessary to verify the effectiveness of the simulation by weight aqueous solution of sodium nitrate (NaNO3). The abrasive
analysis and confirm the choice of optimal tool by means of comparison tool rotated continuously with the spindle at a speed of 1000 rev min−1
tests. [23].

3. Experimental details 3.2. Experimental procedures

3.1. Experimental setup First, the maximum feed rate was explored for the four abrasive
tools via a series of comparison tests, with the aim of obtaining a higher
Fig. 7 shows the experimental setup constructed for this study of machining speed in the rough machining stage and thereby determining
ECMG. A four-axis machine tool was utilized to perform the motion the optimal number of rows of tool-sidewall outlet holes. A straight slot
mode of end milling, defined in terms of linear axes X, Y, Z and a rotary was then machined using the abrasive tool with this optimal number of
axis C. The workpiece was clamped on a specially designed worktable rows. The effects of processing parameters, including applied voltage,
that allowed conduction of electric current. The abrasive tool was in- electrolyte pressure, and feed rate, on the machining performance were
stalled on the bottom of a hollow spindle that allowed a radial supply of analyzed by experiments comparing the MRR, average slot width, and
electrolyte. The motions of the workpiece and the tool were driven and sidewall flatness for different parameter combinations.

13
S. Niu, et al. Journal of Manufacturing Processes 41 (2019) 10–22

Fig. 6. Current density distributions on the machining surface with different tool substrates.

The Inconel 718 workpieces were ultrasonically cleaned and 4. Influence of the number of rows of tool-sidewall outlet holes
weighed before and after each experiment, using an analytical balance
with a precision of 0.01 g. The volumetric MRR was calculated based on To improve processing efficiency, a series of tests were carried out
differential mass measurements and the processing time. Cross-sec- to study the effect of the number of rows of tool-sidewall outlet holes on
tional images of the machined slot were captured by an optical mi- the maximum feed rate at a depth of cut of 10 mm. Comparison tests at
croscope (DVM5000, Leica, Germany). The cross-sectional profile of the applied voltages of 15, 20, 25, and 30 V and electrolyte pressures of 0.2
machined slot was measured by a coordinate measuring machine and 0.6 MPa were conducted. The maximum feed rate was defined as
(ZEISS CONTURA, Germany). In addition, a scanning electron micro- the maximum value of the feed rate at which no spark or short circuit
scope (SEM: S-4800, Hitachi, Japan) was used to observe the surface of occurred during the process with a machining time of 5 min. Each test
the slot sidewall. The surface roughness of the slot sidewall was mea- started with a low feed rate, which was then gradually increased in
sured by a 3D scanning laser microscope (OLS4100, Olympus, Japan). increments of 0.1 mm min−1 until it reached the maximum feed rate.

Fig. 7. The four-axis machine tool used in the experiments.

14
S. Niu, et al. Journal of Manufacturing Processes 41 (2019) 10–22

Fig. 8. Photograph of the four abrasive tools.


Fig. 9. MRR for different experimental parameters.
Table 1
Maximum feed rate of ECMG with different tools. value obtained for the MRR is significantly larger than the MRR in
Group Applied Electrolyte Maximum feed rate (mm min−1) previous work [23,24]. This clearly indicates that the MRR of ECMG
voltage (V) pressure (MPa) can be enhanced dramatically using the improved abrasive tool.
Tool A Tool B Tool C Tool D A cross section at a distance of 10 mm from the slot entrance was
1 15 0.2 0.7 0.9 1.1 1.1
chosen to observe the profile of the machined sidewall. As shown in
2 15 0.6 1.2 1.4 1.6 1.6 Fig. 10, the sidewall profile is inhomogeneous, and visible depressions
3 20 0.2 0.9 1.1 1.4 1.4 and bulges are observed. According to the simulation results for the
4 20 0.6 1.4 1.6 1.9 1.9 electric field, the dissolution rate on the machining surface immediately
5 25 0.2 1 1.3 1.6 1.6
before each row of tool-sidewall outlet holes is relatively low, resulting
6 25 0.6 1.6 1.9 2.2 2.1
7 30 0.2 1 1.4 1.7 1.7 in the bulges on the slot sidewall. In addition, because of the rotation of
8 30 0.6 1.8 2.2 2.5 2.4 the tool in ECMG, the electrolyte may be sprayed directly onto the
machined sidewall, which leads to stray current attacks. As a result,
material removal inevitably continues on the machined area, which
Each test was performed three times with the aim of ensuring repeat- further worsens the uniformity of the sidewall profile. This process is
ability. known as stray removal [30,31]. Obviously, the higher the electrolyte
The test results are listed in Table 1. Among the first three tools, the flow velocity, the higher is the dissolution rate and the greater is the
maximum feed rate in each group of tests was highest for tool C and stray removal, resulting in the depressions on the slot sidewall. Hence, a
lowest for tool A. This is primarily because the increased number of flow field simulation using tool C needs to be performed to study the
rows of outlet holes provides a greater and more uniform electrolyte effects of the arrangement of outlet holes on the flow velocity dis-
flow, which rinses away machining by-products from the frontal gap tribution near the slot sidewall.
more rapidly. Nevertheless, this trend is not always helpful, and the test Fig. 11 shows the flow velocity distribution at section B for different
showed that the maximum feed rate of tool D was not increased in angles of rotation. When the rotating zone has rotated 90°, there are
comparison with that of tool C, and even decreased at higher applied four outlet holes directly opposite the slot sidewall. It can be seen from
voltage and electrolyte pressure. A likely explanation is that with tool Fig. 11a that the flow velocity in the vicinity of each outlet hole is very
D, there are not only more regions with low flow velocity in the front high at section B, and all zones with high flow velocity are arranged
gap but also more regions with low current density in the vertical di- vertically in a line. When the rotating zone has rotated 180° (Fig. 11b),
rection of the machining surface, which leads to a deterioration in the although there is no outlet hole directly facing the slot sidewall, all
uniformity of the dissolution rate distribution on the machining surface. regions with higher flow velocity are arranged in two vertical columns
These test results are in clear agreement with the simulation results. at section B. The simulations indicate that the regions with high flow
Therefore, to obtain a high feed rate, tool C is the optimal choice. velocity are always lined up vertically on the slot sidewall, even when
Furthermore, a previous study has shown that an appropriate high tool C rotates. Moreover, it is also found that the areas of high flow
applied voltage and electrolyte pressure contribute significantly to en- velocity located between two adjacent outlet holes can become en-
hancing electrochemical dissolution in ECMG [23]. As a result, with larged owing to the confluence of two jet flows. Obviously, stray re-
increasing applied voltage and electrolyte pressure, the maximum feed moval is greater at the locations on the machined sidewall with higher
rate of tool C was improved from 1.1 mm min−1 to 2.5 mm min−1. flow velocity. Therefore, it is not surprising that the final sidewall
Tool C was then used to machine a slot at a depth of cut of 10 mm. profile machined by tool C is very uneven.
Experiments were performed with applied voltages of 15, 20, 25, and The effects of the experimental parameters on the machining ac-
30 V and electrolyte pressures of 0.2 and 0.6 MPa. To ensure stability of curacy of the machined sidewall were investigated by determining the
the early processing stage, each experiment started at a low feed rate average slot width and sidewall flatness. The cross-sectional profile of
and a machining length of 5.2 mm, with the feed rate then being in- the slot was measured by the coordinate measuring machine with the
creased to its maximum value at a machining length of 19.8 mm. Each detection scheme shown in Fig. 12. The average slot width was taken as
experiment was repeated three times. the average value of the horizontal distances of all measured points on
Fig. 9 shows the MRR obtained with different experimental para- the two sidewalls located between lines 1 and 3. In addition, the flat-
meters. By increasing the applied voltage and electrolyte pressure, a ness of each sidewall was taken as the difference between the maximum
high feed rate can be applied in machining the slot, which shortens the and minimum horizontal distances of all measured points on the side-
processing time. As a result, with an increase in the maximum feed rate wall located between lines 1 and 2. The average sidewall flatness of
from 1.1 mm min−1 to 2.5 mm min−1, the MRR improves from each slot was then taken as the average of the flatnesses of the left and
111.7 mm3 min−1 to 216.6 mm3 min−1. In this study, even the lowest right sidewalls, which is thus given by

15
S. Niu, et al. Journal of Manufacturing Processes 41 (2019) 10–22

Fig. 10. Cross-sectional profiles of slots machined using tool C with different experimental parameters.

(L1M − L1) + (L2M − L2)


Fs = ,
2 (7)

where L1M and L1 are respectively the maximum and minimum hor-
izontal distances from the left-side measurements, and L2M and L2 are
those from the right-side measurements (see Fig. 12).
Fig. 13a illustrates the effects of the experimental parameters on the
average slot width. Because of the occurrence of stray machining, the
average depth of each machined slot is often larger than the outer
diameter of the abrasive tool. When the applied voltage is fixed, the
average slot width is smaller at a higher electrolyte pressure. A likely
explanation is that the stray machining time can be shortened by using
a larger maximum feed rate at a higher electrolyte pressure, resulting in
a reduced amount of stray removal on the slot sidewall. Furthermore, at
a fixed value of the electrolyte pressure, the average slot width in-
creases with increasing applied voltage. This is mainly because, with Fig. 12. Schematic representation of the machined sidewall measurements.
the increase in applied voltage, although the maximum feed rate has
increased, the stray current has gone up even more per unit time, re-
sulting in a greater amount of stray removal on the slot sidewall. distribution of stray removal across the sidewall profile, resulting in a
Therefore, the average slot width increases monotonically as the ap- greater average sidewall flatness. In the rough stage of ECMG, a higher
plied voltage increases within the range from 15 V to 30 V. applied voltage should be considered to achieve higher machining ef-
Fig. 13b shows the results for the average sidewall flatness with ficiency. Hence, the application of a larger electrolyte pressure not only
different experimental parameters. It can be seen that the average further increases the maximum feed rate and MRR but also noticeably
sidewall flatness increases with increasing applied voltage at a fixed improves the flatness of the machined sidewall.
electrolyte pressure, whereas it becomes smaller at a higher electrolyte
pressure for a fixed applied voltage. According to the results of the flow 5. Influence of the mode of arrangement of the tool-sidewall
field simulation, the middle part of the machined sidewall is subject to a outlet holes
higher flow velocity of electrolyte than the ends, leading to a greater
amount of stray removal in the middle part. Obviously, a high applied Even though the optimal number of rows of tool-sidewall outlet
voltage and a low electrolyte pressure both lead to a seriously uneven holes has been determined and the results of machining a slot using the

Fig. 11. Flow velocity distribution at section B with tool C at different angles of rotation.

16
S. Niu, et al. Journal of Manufacturing Processes 41 (2019) 10–22

Fig. 13. Average slot width and average sidewall flatness for different values of the experimental parameters.
(a) Schematic of tool substrates.
(b) Photograph of abrasive tools.

optimal abrasive tool have been shown, a few noticeable defects are tool E, the outlet holes are arranged in a spiral manner, rather than the
observed for tools with vertically arranged outlet holes. For example, vertical arrangement in the reference case (tool C). In terms of the
with tool C, there is a vertical distribution of regions of high flow ve- design parameters of the two tool substrates, compared with tool C, the
locity on the slot sidewall, resulting in the formation of an uneven only difference in tool E is that the hole centers are displaced by an
profile of the machined sidewall. Based on the results of previous work, angle of 15° between any two adjacent rows. The two abrasive tools
the mode of arrangement of the outlet holes can be adjusted with the (Fig. 14b) have the same diamond particle size and concentration.
aim of further improving machining performance.

5.1. Arrangement of tool-sidewall outlet holes 5.2. Simulation results

The influence of the arrangement of the tool-sidewall outlet holes is An electric field simulation using tool E was carried out to obtain
examined using the tool substrates shown schematically in Fig. 14a. In the current density distribution on the machining surface. It can be seen
found from Fig. 15 that the regions of low current density are arranged
in a spiral manner corresponding to the positions of the tool-sidewall
outlet holes. This means that the spiral arrangement of the outlet holes
prevents excessive concentration of regions with slow dissolution rate
in a vertical direction on the machining surface. Thus, the use of tool E
improves the uniformity of the electric field distribution on the ma-
chining surface compared with tool C.
A flow field simulation was also performed for tool E, and the flow
velocity distributions at section B for different angles of rotation are
shown in Fig. 16. When the rotating zone has rotated 90°, there is only
an outlet hole in the first row directly facing the slot sidewall. It can be
seen from Fig. 16a that the regions of high flow velocity are distributed
in a slanting manner instead of with a vertical alignment. When the
rotating zone has rotated 180°, there is only an outlet hole in the third
row directly facing the slot sidewall. It can be seen from Fig. 16b that

Fig. 15. Current density distribution on the machining surface when tool E is
Fig. 14. Different arrangements of tool-sidewall outlet holes. used.

17
S. Niu, et al. Journal of Manufacturing Processes 41 (2019) 10–22

Fig. 16. Flow velocity distribution at section B with tool E at different angles of rotation.

the distribution of the regions of high flow velocity is still dispersed and
tilted, rather than being concentrated in a vertical direction. Further-
more, it can be noted from the two velocity contours that there are
fewer areas in which there is repeatedly a high flow velocity. This
means that a spiral arrangement of outlet holes can lead to better
mixing through tool rotation compared with a vertical arrangement.
Hence, the use of tool E also improves the uniformity of the flow field
distribution on the machined sidewall compared with tool C.

5.3. Comparison of machining performance

In this subsection, the machining performance of tool E is compared


with that of tool C. First, the maximum feed rate was tested at a depth
of cut of 10 mm using applied voltages of 15, 20, 25, and 30 V and an
electrolyte pressure of 0.6 MPa. The test procedure was the same as
before, and the test results are listed in Table 2. It was found that the
maximum feed rate using tool E was always higher than that with tool Fig. 17. MRR of slot machined using tools C and E.
C. This is mainly because tool E provides a more uniform electric field
than tool C on the machining surface of the workpiece. The test results
homogeneous stray removal along the vertical direction of the sidewall.
are consistent with those of the simulation.
These experimental results are in good agreement with those of the flow
A slot was then machined at a depth of cut of 10 mm using tool E.
field simulations.
Experiments were carried out at applied voltages of 15, 20, 25, and 30 V
The average slot width obtained with tools C and E is plotted in
and an electrolyte pressure of 0.6 MPa, with the same experimental
Fig. 19a. For the two abrasive tools, the stray current markedly in-
procedure as before. Fig. 17 shows the results for MRR with the two
creases per unit time with a rise in the applied voltage, and so do the
abrasive tools. It can be seen that the MRR achieved with tool E is
amount of stray removal and the average slot width. In addition, it is
always greater than that with tool C. This is because the maximum feed
worth noting that the average slot width for tool E is always smaller
rate of tool E in every group of experiments is relatively large, resulting
than that for tool C. A likely explanation is that the sidewall machined
in a reduced machining time and increased MRR. Furthermore, owing
with tool E is subject to a shorter stray machining time, and thus to a
to the fact that the maximum feed rate and the amount of stray removal
relatively small amount of stray removal, than that machined with tool
increase with a rise in applied voltage, the MRR of machining the slot
C.
using tool E is also improved from 150 mm3 min−1 to 234.6 mm3
The results of the average sidewall flatness for the two abrasive tools
min−1.
are shown in Fig. 19b. It can be seen that the average sidewall flatness
The cross-sectional profiles of the slots machined by the two abra-
with tool E is always lower than that with tool C. Moreover, for an
sive tools are shown in Fig. 18. It can be seen that the sidewall profile
applied voltage greater than 15 V, compared with the case of tool C, the
machined using tool E is flatter than that machined using tool C. The
average sidewall flatness obtained with tool E increases only slightly.
reason for this is that tool E provides a more uniform flow field dis-
This indicates that tool E is more appropriate for the stage of rough
tribution than tool C on the machined sidewall, leading to relatively
machining with high applied voltage than tool C. With an applied
voltage of 30 V, the average sidewall flatness produced by tool C is
Table 2
549.6 μm, but it is only 340.5 μm for tool E. Hence, it is clear that the
Maximum feed rates of ECMG with tools C and E.
use of tool E can effectively decrease the allowance in the machined
Group Applied voltage Electrolyte pressure Maximum feed rate (mm sidewall for subsequent finishing.
(V) (MPa) min−1)
Fig. 20 shows images of the sidewalls produced by the two abrasive
Tool C Tool E tools at an applied voltage of 30 V. It can be seen that many deep pits
are present on the sidewall machined with tool C (Fig. 20a), resulting in
1 15 0.6 1.6 1.7 a honeycomb pit structure。In the stage of rough machining with
2 20 0.6 1.9 2
ECMG, most of the current is constrained within the machining region,
3 25 0.6 2.2 2.4
4 30 0.6 2.5 2.7 whereas the current density drops dramatically outside this region. As a
result, the sidewall surface of the machined slot is exposed to low-

18
S. Niu, et al. Journal of Manufacturing Processes 41 (2019) 10–22

Fig. 18. Cross-sectional profiles of slots machined using tools C and E.


(a) Average slot width (b) Average sidewall flatness

Fig. 19. Average slot width and average sidewall flatness with tools C and E.

current-density dissolution, which leads to selective corrosion of the mainly by peak–valley configurations formed through pitting corrosion.
material or pitting corrosion. Additionally, it is worth noting that the The peak–valley distance with tool C is noticeably larger than that with
pitting effect is reduced on the sidewall machined with tool E tool E. As a result, the surface roughness with tool C is Ra = 2.509 μm,
(Fig. 20b). The flow field may be the main reason for this occurrence, whereas it is only Ra = 1.65 μm with tool E. This indicates that, com-
because the regions of high electrolyte flow velocity are more widely pared with tool C, tool E again significantly improves the surface
dispersed on the machined sidewall when tool E is used, thereby alle- quality of the machined sidewall in ECMG.
viating the local concentrations of pitting corrosion found with tool C.
The surface roughnesses of the sidewalls machined at an applied
voltage of 30 V were measured, as shown in Fig. 21. The surface 6. Refining the machined sidewall with ECMG
morphologies of the sidewalls machined by both tools are characterized
It should be noted that although the machining efficiency in the

Fig. 20. Surface topographies of the sidewalls machined with tools C and E, respectively, at an applied voltage of 30 V.

19
S. Niu, et al. Journal of Manufacturing Processes 41 (2019) 10–22

Fig. 21. Surface roughness of the machined sidewall at an applied voltage of 30 V with tools C and E.

rough machining stage is high, the profile of the machined sidewall is


quite uneven and its surface is rough. Even with the use of tool E, the
flatness and surface roughness of the slot sidewall are still insufficient to
meet the demands of finishing. Hence, after rough machining has been
completed, the machined sidewall needs to be further refined. With
ECMG, it is possible to do this simply by setting suitable finishing
parameters, without any need to replace the tool.
In the stage of rough machining with ECMG, efficient removal of the
large machining allowance for the slot was achieved using tool E at an
applied voltage of 30 V. Obviously, for the finishing operation, only a
small amount of material needs to be removed from the slot sidewall.
The anodic polarization curves of Inconel 718 are plotted in Fig. 22.
Once the potential has exceeded a value of about 1.3 V, the current
density increases gradually, indicating that electrochemical dissolution
is occurring on the anode surface. Thus, a low applied voltage, small
depth of cut, low electrolyte pressure, and high feed rate should be Fig. 22. Polarization curves of Inconel 718 in 10 wt.% NaNO3 electrolyte at
30 °C.
considered for finish machining with ECMG, with the aim of reducing
(a) Cross-sectional photographs.
stray current attack and tool wear.
(b) Cross-sectional profiles.
The following finishing parameters were adopted in this stage: an
applied voltage of 2 V, a depth of cut of 0.03 mm, an electrolyte pres-
sure of 0.2 MPa, and a feed rate of 25 mm min−1. The other machining smearing, and redeposited chips, which are usually considered to be
parameters were the same as before, and a specified finishing allowance microstructural features of the ground surface generation of Inconel
could be attained by using multiple passes. Cross-sectional profiles of 718 [32]. The experimental results indicate that the stray-corrosion
slots obtained by both rough and finish machining with ECMG are effects due to rough machining can be eliminated effectively by finish
shown in Fig. 23. It can be seen that the sidewall profile is very flat after machining. In addition, Fig. 25 shows the ground surface morphology
finish machining: the average sidewall flatness is 69.5 μm and has a of the sidewall after finish machining with ECMG. It can be seen that
79.6% improvement compared with rough machining. the peak–valley distance after finish machining is significantly less than
Fig. 24 shows the surface topography of the machined sidewalls in that after rough machining. The ground surface roughness is Ra =
the finishing stage of ECMG. It is interesting to note that there is vir- 0.648 μm and has a 60.7% improvement compared with the rough
tually no pitting corrosion on the sidewall surface. Furthermore, some machined surface.
surface defects can be observed, such as wide and deep striations, The significant improvements in the flatness and surface quality of

20
S. Niu, et al. Journal of Manufacturing Processes 41 (2019) 10–22

the slot sidewall confirm that ECMG with the same abrasive tool can be
applied for successive rough and finish machining. The results of this
work provide evidence to promote the further development of ECMG
technology and its wider application in the aircraft manufacturing in-
dustry.

7. Conclusions

Abrasive tools with different designs of tool-sidewall outlet holes


have been employed for ECMG of Inconel 718 with the aim of im-
proving machining performance. The required depth of cut was set at
10 mm in the rough machining stage. A selection of tools were designed
with two, three, four and five rows of tool-sidewall outlet holes and
with different layouts of these holes, either vertical or in a spiral ar-
rangement. Based on the results of both simulations and experiments,
the following conclusions can be drawn:

(1) The flow field simulations indicate that with increasing number of
rows of tool-sidewall outlet holes, the uniformity of the flow field at
the front gap is at first enhanced but then deteriorates. The electric
field simulations reveal that a greater number of rows also leads to
poorer uniformity of the electric field in the vertical direction of the
machining surface. The test results show that a higher maximum
feed rate can be obtained by using an abrasive tool with four rows
of outlet holes.
(2) The experimental results on machining a slot using the optimal
abrasive tool show that the MRR is increased at higher applied
voltage, electrolyte pressure and feed rate. Furthermore, it is found
Fig. 23. Cross-sectional photographs and profiles of slots produced by rough that the average slot width and average sidewall flatness both in-
machining and finish machining with ECMG. crease with increasing applied voltage and feed rate at a fixed
electrolyte pressure, but they become smaller at a higher electrolyte
pressure and feed rate for a fixed applied voltage.
(3) A major defect of the abrasive tool with a vertical alignment of tool-
sidewall outlet holes is found to be the formation of an uneven
profile on the machined sidewall. To enhance the machining per-
formance, an abrasive tool with a spiral arrangement of outlet holes
has been proposed. The experimental results with this tool show
that its use noticeably improves the flatness of the machined side-
wall.
(4) The uneven sidewall profile and rough sidewall surface generated
by the rough machining stage can be dramatically modified in a
finishing stage using ECMG with suitable parameters. Compared
with that obtained from rough machining, the average sidewall
flatness decreases from 340.5 μm to 69.5 μm and the sidewall sur-
face roughness decreases from 1.65 μm to 0.648 μm after finish
machining.

Fig. 24. Surface topography of the sidewall after the finishing stage with Acknowledgments
ECMG.
The authors wish to acknowledge financial support from the

Fig. 25. Surface roughness of the machined sidewall after finish machining with ECMG.

21
S. Niu, et al. Journal of Manufacturing Processes 41 (2019) 10–22

National Key Research and Development Program of China 2018;140(7):071009.


(2018YFB1105902) and Funding of Jiangsu Innovation Program for [16] Zaborski S, Lupak M, Poros D. Wear of cathode in abrasive electrochemical grinding
of hardly machined materials. J Mater Process Technol 2004;149(1–3):414–8.
Graduate Education (Grant KYLX16_0316). [17] Goswami RN, Mitra S, Sarkar S. Experimental investigation on electrochemical
grinding (ECG) of alumina-aluminum interpenetrating phase composite. Int J Adv
References Manuf Technol 2009;40(7–8):729–41.
[18] Tehrani AF, Atkinson J. Overcut in pulsed electrochemical grinding. Proc Inst Mech
Eng B J Eng Manuf 2000;214(4):259–69.
[1] Pan ZP, Feng YX, Hung TP, Jiang YC, Hsu FC, Wu LT, et al. Heat affected zone in the [19] Mogilnikov VA, Chmir MY, Timofeev YS, Poluyanov VS. Diamond-ECM grinding of
laser-assisted milling of Inconel 718. J Manuf Process 2017;30:141–7. sintered hard alloys of WC–Ni. Procedia CIRP 2016;42:143–8.
[2] Fang ZL, Obikawa T. Turning of Inconel 718 using inserts with cooling channels [20] Roy S, Bhattacharyya A, Banerjee S. Analysis of effect of voltage on surface texture
under high pressure jet coolant assistance. J Mater Process Technol in electrochemical grinding by autocorrelation function. Tribol Int
2017;247:19–28. 2007;40(9):1387–93.
[3] Zhu ZW, Wang DY, Bao J, Wang NF, Zhu D. Cathode design and experimental study [21] Hascalık A, Caydas U. A comparative study of surface integrity of Ti–6Al–4V alloy
on the rotate-print electrochemical machining of revolving parts. Int J Adv Manuf machined by EDM and AECG. J Mater Process Technol 2007;190(1–3):173–80.
Technol 2015;80(9–12). 1957–63. [22] Qu NS, Zhang QL, Fang XL, Ye EK, Zhu D. Experimental investigation on electro-
[4] Ucak N, Cicek A. The effects of cutting conditions on cutting temperature and hole chemical grinding of Inconel 718. Procedia CIRP 2015;35:16–9.
quality in drilling of Inconel 718 using solid carbide drills. J Manuf Process [23] Li HS, Niu S, Zhang QL, Fu SX, Qu NS. Investigation of material removal in inner-jet
2018;31:662–73. electrochemical grinding of GH4169 alloy. Sci Rep 2017;7:3482.
[5] Sinha MK, Setti D, Ghosh S, Rao PV. An investigation on surface burn during [24] Li HS, Fu SX, Zhang QL, Niu S, Qu NS. Simulation and experimental investigation of
grinding of Inconel 718. J Manuf Process 2016;21:124–33. inner-jet electrochemical grinding of GH4169 alloy. Chin J Aeronaut
[6] Yilmaz B, Karabulut S, Gullu A. Performance analysis of new external chip breaker 2018;31(3):608–16.
for efficient machining of Inconel 718 and optimization of the cutting parameters. J [25] Niu S, Qu NS, Li HS. Investigation of electrochemical mill-grinding using abrasive
Manuf Process 2018;32:553–63. tools with bottom insulation. Int J Adv Manuf Technol 2018;97(1–4):1371–82.
[7] Zhang JC, Xu ZY, Zhu D, Su WF, Zhu D. Study of tool trajectory in blisk channel [26] Niu S, Qu NS, Fu SX, Fang XL, Li HS. Investigation of inner-jet electrochemical
ECM with spiral feeding. Mater Manuf Process 2017;32(3):333–8. milling of nickel-based alloy GH4169/Inconel 718. Int J Adv Manuf Technol
[8] Zhou X, Zhang DH, Luo M, Wu BH. Chatter stability prediction in four-axis milling 2017;93(5–8):2123–32.
of aero-engine casings with bull-nose end mill. Chin J Aeronaut [27] Zhu D, Zeng YB, Xu ZY, Zhang XY. Precision machining of small holes by the hybrid
2015;28(6):1766–73. process of electrochemical removal and grinding. CIRP Ann Manuf Technol
[9] Lupak M, Zaborski S. Simulation of energy consumption in electrochemical grinding 2011;60:247–50.
of hard-to-machine materials. J Appl Electrochem 2009;39(1):101–6. [28] Li S, Wu Y, Yamamura K, Nomura M, Fujii T. Improving the grindability of titanium
[10] Mohammad A, Wang DW. Electrochemical mechanical polishing technology: recent alloy Ti–6Al–4V with the assistance of ultrasonic vibration and plasma electrolytic
developments and future research and industrial needs. Int J Adv Manuf Technol oxidation. CIRP Ann Manuf Technol 2017;66(1):345–8.
2016;86(5–8):1909–24. [29] Mishra P, Ein-Mozaffari F. Using computational fluid dynamics to analyze the
[11] Maksoud TMA, Brooks AJ. Electrochemical grinding of ceramic form tooling. J performance of the Maxblend impeller in solid-liquid mixing operations. Int J
Mater Process Technol 1995;55(2):70–5. Multiph Flow 2017;91:194–207.
[12] Puri AB, Banerjee S. Multiple-response optimisation of electrochemical grinding [30] Wang DY, Zhu ZW, Bao J, Zhu D. Reduction of stray corrosion by using iron coating
characteristics through response surface methodology. Int J Adv Manuf Technol in NaNO3 solution during electrochemical machining. Int J Adv Manuf Technol
2013;64(5–8):715–25. 2015;76(5–8):1365–70.
[13] Kozak J, Oczos KE. Selected problems of abrasive hybrid machining. J Mater [31] Wang XD, Qu NS, Fang XL. Reducing stray corrosion in jet electrochemical milling
Process Technol 2001;109(3):360–6. by adjusting the jet shape. J Mater Process Technol 2019;264:240–8.
[14] Lauwers B, Klocke F, Klink A, Tekkaya AE, Neugebauer R, McIntosh D. Hybrid [32] Qian N, Ding WF, Zhu YJ. Comparative investigation on grindability of K4125 and
processes in manufacturing. CIRP Ann Manuf Technol 2014;63(2):561–83. Inconel 718 nickel-based superalloys. Int J Adv Manuf Technol
[15] Li SS, Wu YB, Nomura M, Fujii T. Fundamental machining characteristics of ul- 2018;97(5–8):1649–61.
trasonic-assisted electrochemical grinding of Ti-6Al-4V. ASME J Manuf Sci Eng

22

You might also like