You are on page 1of 11

Journal of Materials Processing Tech.

311 (2023) 117807

Contents lists available at ScienceDirect

Journal of Materials Processing Tech.


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

Adapting ‘tool’ size using flow focusing: A new technique for


electrochemical jet machining
Ivan Bisterov a, Sidahmed Abayzeed b, Alistair Speidel a, Mirco Magnini c, Mohamed Zubayr a,
Adam T. Clare a, *
a
Advanced Manufacturing Research Group, University of Nottingham, Nottingham NG7 2RD, United Kingdom
b
Optics and Photonics Research Group, University of Nottingham, Nottingham NG7 2RD, United Kingdom
c
Fluids and Thermal Engineering Research Group, University of Nottingham, Nottingham NG7 2RD, United Kingdom

A R T I C L E I N F O A B S T R A C T

Associate Editor: Jingjing Li Electrochemical jet manufacturing methods exploit the localised interaction of an electrified jet with a
conductive workpiece. This has been exploited by numerous researchers to deposit and remove material. In
Keywords: recent studies, the same approach has been used as an analysis tool to measure and evaluate engineering
Electrochemical jet machining components. The capability of this manufacturing method is limited by the resolution, which is governed by the
Electrochemical machining
diameter of the jet. Although approaches have been taken to reduce the kerf or interaction volume in the process,
Precision machining
it is the jet diameter still provides the fundamental limit. In this study, a new method is proposed which takes
Flow focusing
Tool design advantage of a constriction effect to reduce the jet diameter through flow focusing, which occurs in coaxial two-
phase flows. A novel nozzle arrangement is presented which demonstrates jets can be constricted by 79% leading
to 54% reduction in machined kerf width. The limitations of this method are investigated in the context of fluid
dynamic constraints, identifying optimum operating regions to utilise the approach in a computer numerical
control machine tool arrangement. This enables continuously varying tool size in-process, which is analogous to
other energy beam processes where spot size can be adjusted with a corresponding response influence.

1. Introduction liner material through application of surface features which acted as


lubricant reservoirs. More recently, the creation of anti-bacterial sur­
Electrochemical micromachining methods, reviewed in detail by faces through EJM was reported by Lutey et al. (2021). In addition to
Saxena et al. (2018), offer the capability to process metallic materials material removal, by reversing the polarity of the power supply, the
independent of their hardness or mechanical properties and without any electrochemical jet apparatus can be used to perform localised electro­
thermally induced microstructural changes. This presents an advantage deposition to build complicated 3-dimensional structures as demon­
compared to other machining processes such as milling and electrical strated by Wei et al. (2022). Bisterov et al. (2022) used the
discharge machining. Electrochemical jet machining (EJM) is an elec­ electrochemical jet as a measurement tool utilising the fact that its
trochemical micromachining process where the electrolyte is delivered electrical resistance can be correlated to the distance between the nozzle
through a nozzle cathode to generate a jet impinging onto the anodic and the surface.
work piece. Under the influence of applied voltage between the nozzle One of the key limitations restricting wider application of electro­
and work piece, material is locally removed from the impingement re­ chemical jet methods is the spatial resolution achievable for localised
gion allowing the fabrication of surface features to impart functional material processing as identified in a recent review by Speidel et al.
properties to engineering components. Yang et al. (2016) applied EJM to (2022). Generally, this is defined by jet diameter controlled by the in­
manufacture dimples on a superhydrophobic aluminium alloy surface ternal diameter of the nozzle electrode. Improvements in resolution are
and demonstrated control over the adhesion force to the surface hence limited by the ability to manufacture small nozzles and deliver
dependent on the machining parameters. Walker et al. (2019) demon­ controllable electrolyte flow through them. Li et al. (2021) proposed an
strated 37% reduction in coefficient of sliding friction in Al-Si cylinder in-situ method for nozzle preparation through metal electrodeposition

* Corresponding author.
E-mail address: adam.clare@nottingham.ac.uk (A.T. Clare).

https://doi.org/10.1016/j.jmatprotec.2022.117807
Received 22 July 2022; Received in revised form 6 October 2022; Accepted 22 October 2022
Available online 26 October 2022
0924-0136/© 2022 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
I. Bisterov et al. Journal of Materials Processing Tech. 311 (2023) 117807

Fig. 1. Flow focusing configuration for EJM showing the focusing air nozzle assembled coaxially with the nozzle electrode. The pressure differential across the
focusing orifice constricts the jet to diameters smaller than the nozzle electrode internal diameter. An additional outer air passage is provided to prevent electrolyte
accumulation on the work piece surface. Material removal occurs under the influence of the applied voltage between the nozzle and the work piece.

on its internal surface from a copper sulphate solution. Machining of 20 2016) to guide the focusing gas stream have been reported and jet di­
µm wide and 5 µm deep microgrooves was demonstrated, which to the ameters of the order of 300 nm have been achieved (Deponte et al.,
authors’ knowledge, represents the narrowest kerf machined by EJM 2011). In addition to the standard gas-liquid flow focusing, Gañán-Calvo
reported in literature. The use of nozzles of this size however could pose et al. (2007) developed a multi-stage focusing arrangement consisting of
challenges to the design of the electrolyte filtration system to prevent a gas stream focusing a viscous sheath liquid which in turn focuses a core
clogging and maintain reliable operation. As such the practical appli­ liquid stream, with the objective of improving stability and reducing the
cation could be challenging. Clare et al. (2018) showed improvement of minimum jet diameter. A fundamental modification of the flow focusing
machining resolution for a fixed nozzle size through the combination of principle resulted from the use of a sharp hypodermic needle to inject
nozzle profile modification, nozzle inclination and addition of dopants the focused liquid into the focusing fluid steam with the needle tip
to the electrolyte, resulting in machined kerf of the order of nozzle size. positioned along the centre line of the focusing orifice (Acero et al.,
Saxena et al. (2020) integrated a laser beam co-axially with the jet which 2013) – this produced a Couette-type flow leading to ejection of a liquid
also allowed constriction of material removal by locally increasing the ligament from the sharp tip. Lower stable flow rates than the standard
temperature and consequently the conductivity of the electrolyte. Wu configuration were reported achieving submicrometric jets emanating
et al. (2020) fabricated features significantly smaller than nozzle from a 160 µm internal diameter needle. Currently, the key application
diameter through the photolithographic application of masks to the of flow focusing methods is sample delivery for serial femtosecond X-ray
work piece surface, however this adds an extra step in the process it is crystallography (Chapman et al., 2011) for structure determination of
not applicable to the measurement mode of the apparatus. Beyond the biochemical molecules where the destruction of the crystal during
limitations in terms of resolution, manual nozzle electrode changes are measurement necessitates continuous replenishment of the sample.
required in EJM if features of different scales are to be produced on the Other proposed applications areas include polymer fibre spinning
same work piece. This results in non-value-added time and can incur (Ponce-Torres et al., 2019), fabrication of microparticles for drug de­
significant costs in production. livery (Keohane et al., 2014) and deposition of polymer films (Ponce-­
In this work, a novel tooling concept for electrochemical jet Torres et al., 2017). The compressed air-film encircling jet
machining based on the hydrodynamic principle of flow focusing is electrodeposition reported by Xinchao et al. (2020) is effectively an
investigated, which addresses the limitations of machining resolution application of the flow focusing principle to electrochemical jet depo­
and fixed nozzle sizes. The flow focusing method, explored previously in sition. The addition of a focusing orifice downstream of the jet resulted
fundamental bench top fluid dynamics experiments by Gañán-Calvo in reduction of copper deposit diameter of approximately 23% from
(1998), is capable of producing micrometric liquid jets and droplets 1.32 mm to 1.01 mm produced with a 1 mm nozzle electrode.
without the need for small nozzles. In the original configuration of this The proposed tooling design in this work consists of an additively
approach, the tip of the nozzle generating the jet is positioned close to an manufactured focusing air nozzle assembled coaxially with the nozzle
orifice and a co-flowing gas stream is applied, so that both the liquid jet electrode. In addition to the focusing air stream (see Fig. 1) a second
and the gas flow through the orifice. At sufficiently low liquid flow rates, pathway is provided surrounding the inner nozzle to deliver compressed
the pressure of the gas stream stabilises a liquid meniscus at the nozzle air for clearing electrolyte from the work piece surface. This prevents
tip from which a jet much smaller than the orifice size is emitted (see electrolyte accumulation irrespective of the orientation of the jet with
Fig. 1). Vega et al. (2010) showed that he geometric parameters of the respect to gravity and is an arrangement typically used in EJM (Speidel
system, namely nozzle internal diameter D1 , orifice diameter D and et al., 2022). The supply pressures of the two pathways are controlled
nozzle tip-to-orifice distance H (see Fig. 1) define the stability of the separately by two independent pressure regulators. It is proposed that
meniscus, the minimum achievable flow rate and consequently the through the application of this flow focusing configuration the resolu­
minimum jet diameter for which continuous jetting is possible before the tion of electrochemical jet techniques can be decoupled from nozzle
onset of the so called ‘dripping’ instability. Different embodiments of the diameter, allowing significant reductions of machined kerf without the
flow focusing principle utilising flame-shaped glass capillary tubes need for small nozzle electrodes. According to Gañán-Calvo (1998) for
(Acero et al., 2012), micro-injection moulded nozzles (Beyerlein et al., liquids with sufficiently low viscosity (as high as 0.1 Pa.s for typical
2015) and more recently additively manufactured nozzles (Nelson et al., applied gas pressure differentials across the orifice Δpg > 100 mbar,

2
I. Bisterov et al. Journal of Materials Processing Tech. 311 (2023) 117807

viscous stresses and surface tension stresses are negligible relative to the Table 1
pressure differential Δpg and the focused jet diameter djet is only Machining parameters applied dependent on flow rate.
dependent on Δpg , the liquid flow rate Q and the liquid density ρl : Flow rate Expected diameter Estimated jet Voltage Feed rate
( )1/4 (ml/min) from Eq. (1) (µm) resistance from Eq. (V) (mm/s)
8ρl (2) (kΩ)
djet = Q1/2 (1)
π2 Δpg 7.5 172 6.0 298.8 0.30
5.5 147 7.4 270.1 0.25
The fact that for a fixed liquid density, the jet diameter is controlled 4.0 126 9.4 245.1 0.21
by the flow rate and the pressure drop across the focusing orifice as 3.0 109 11.8 230.5 0.19
2.5 99 13.7 224.2 0.17
expressed by Eq. (1), enables continuously variable tool size by selection
2.0 88 16.6 215.9 0.15
of these parameters without the need for a physical tool change. The 1.5 77 21.4 207.6 0.13
proposed tooling dispenses with the technical challenges associated with 1.0 63 31.0 200.8 0.11
improving the resolution of the conventional EJM configuration such as
the need to fabricate sub-100 µm nozzle electrodes, the risk of clogging
range 100 mbar – 300 mbar and electrolyte flow rates in the range
due to salt crystallisation, and the high electrolyte supply pressures
0.25–2 ml/min. The applied pressures are typical values reported in
required to deliver consistent flow through a small orifice.
fundamental flow focusing studies (Acero et al., 2012). The working and
Other advantages of the flow focusing approach include automated
sense electrode of the potentiostat were connected to the workpiece and
control over machining resolution, reduction of non-value-added time
the counter and reference electrode were connected to the nozzle. Si­
associated with tool changes and overall enhanced process agility. The
nusoidal current waveform of amplitude 100 µA and frequency 10 kHz
objective of this study is to characterise the performance of the devel­
was applied as the excitation signal to reduce the influence of electrode
oped tooling and demonstrate its application to EJM.
overpotentials on the resistance measurement. The nozzle was trans­
lated in the Z-direction between 0.2 mm and 2 mm stand-off distance
2. Materials and methods
while the jet was impinging on the workpiece and the real part of the
electrical impedance was sampled every 10 µm. The initial position
2.1. Tooling design and apparatus
above the workpiece was set with estimated accuracy ± 0.05 mm by
using a feeler gauge and the position during measurement was read
The flow focusing tool was produced by stereolithography using a
directly from the axis encoder.
Formlabs Form 2 3D printer and Formlabs Rigid 4000 photocurable
resin. An orifice with diameter D = 1 mm was drilled in the focusing
nozzle after assembly to ensure alignment with the stainless-steel nozzle 2.3. Optical imaging of the jet
electrode (internal diameter D1 = 200 µm). According to the experi­
ments conducted by Vega et al. (2010) and the derived scaling laws, the The jet was also inspected using an optical microscopy configuration.
optimum distance between the nozzle and orifice for achievement of The jet and the surrounding field were illuminated uniformly using a
1 collimated incoherent light source (Thorlabs M625F2 LED, wavelength:
minimal flow rate is approximately Hopt = 0.6D(D1 /D)4 ≈ 0.41 mm. Due
625 nm). The transmitted light was used to image the jet on CMOS
to manufacturing and assembly tolerances the actual distance in the
camera (Thorlabs DCC1545M, pixel size: 5 µm). The microscope con­
configuration used in this study was H = 0.66 mm ( ± 0.04 mm) with
sisted of a 10x objective lens (Mitutoyo, numerical aperture NA = 0.28,
theoretical minimum stable flow rate of 0.12 ml/min based on the
focal length: 20 mm) and a tube lens (focal distance of 160 mm)
experimental results of Vega et al. (2010) and a corresponding jet
configured as an infinite conjugate system to provide 8x magnification
diameter of 22 µm at Δpg = 150 mbar (Eq. (1)). The thickness of the
where each pixel is equivalent to 0.63 µm.
throat was T = 0.33 mm. The new end effector was integrated within a
The jet was positioned in focus using the axes of the EJM apparatus.
3-axis computer-numerically controlled set-up for EJM described in
Jet diameter was measured near the flow focusing orifice exit directly
detail elsewhere (Mitchell-Smith et al., 2018). To enable precise control
from the acquired images. For the imaging experiments, deionised water
of liquid flow rate, electrolyte was supplied to the nozzle by a KDS 200
was used instead of the electrolyte. According to Eq. (1), the diameter of
syringe pump with a specified flow accuracy of ± 1%. The supplied 1/4
pressure of the focusing air was controlled by means of a Festo VEAB the flow-focused jet is proportional to ρl , so the difference in density
pneumatic regulator with a range of 5 – 1000 mbar and accuracy of between deionised water (≈997 kg/m3) and the electrolyte solution
(1030 kg/m3) employed for machining would result in an error in
± 7.5 mbar. An aqueous solution of sodium nitrate (NaNO3) with a diameter of approximately 0.8%, which is acceptable for the purposes of
nominal concentration of 17 wt% was used throughout this study with this study.
an average conductivity σ = 120.1 mS/cm (standard deviation S.D. =
0.7 mS/cm). To compensate for the variation of conductivity, the exact 2.4. Machining experiments
concentration of the solution was adjusted during the study by addition
of deionised water or salt to maintain the conductivity within ± 1% of Following the characterisation of electrical resistance and optical
the stated value. The conductivity values were acquired using a Mettler imaging, optimisation of the stand-off distance for machining was per­
Toledo S3 conductivity meter with a specified accuracy of ± 0.5%. The formed by machining 12 mm long straight channels while simulta­
density of the employed aqueous solution was 1030 kg/m3, measured neously translating the nozzle in the z-direction at flow rates of 1 ml/
volumetrically at 23 ◦ C. min and 0.5 ml/min and air pressure of 150 mbar. Stainless steel (Grade
304) was chosen as the workpiece material. Machining was performed at
2.2. Electrical resistance measurement controlled current with supplied currents of 8 mA (estimated average
current density based on jet cross-sectional area J ≈251 A/cm2) for the
The electrical resistance response of the impinging electrolyte jet as a 1 ml/min flow rate and 6 mA (J ≈376 A/cm2) for the 0.5 ml/min and
function of stand-off distance is known to be dependent on the jet cross- corresponding horizontal feed rates of 0.11 and 0.07 mm/s respectively
sectional area (Bisterov et al., 2022), so this was acquired to characterise (see Table 1). The motion in the Z-direction was programmed so that the
the behaviour of the flow-focusing tooling electrically. Measurement Z-axis traverses from 1.2 mm to 0.4 mm stand-off distance over 4 mm of
was performed using Ivium Vertex One potentiostat according to the movement in the X-direction. This allowed to evaluate the effect that
methodology described by Bisterov et al. (2022) at air pressures in the stand-off distance has on machining performance and establish the

3
I. Bisterov et al. Journal of Materials Processing Tech. 311 (2023) 117807

Fig. 2. Characterisation of the resistance of flow-focused electrolyte jets. (a) Exemplar resistance response curves at Δp = 150 mbar. (b) Measured average slope from
the linear region of the resistance curve against the optically measured jet diameters. Error bars represent standard deviation from 4 repeated response curves. (c)
Optical images of the jets generated at the conditions in a.

optimum value for further experiments. based on the theoretical jet diameter and resistance. Machining of
Following this, the stand-off distance was fixed at the value of channels was performed with feed rates according to Table 1, which
L= 0.9 mm where stable machining was observed and the influence of were decreased proportionally to the expected jet diameter. This
flow rate variation on kerf width was investigated at a fixed pressure approach was chosen to compensate for the reduced exposure time to
differential of 150 mbar. For comparison of machining performance the jet with reduction of jet diameter and maintain a more uniform
across different jet diameters, it is desirable that a constant current depth of cut across machining conditions. All machining parameters
density is maintained across experimental conditions, since the disso­ used in this experiment as a function of electrolyte flow rate are sum­
lution efficiency of stainless steel in sodium nitrate is known to vary with marised in Table 1.
current density (Datta et al., 1989). The high electrical resistance of the To compare flow focusing against the conventional EJM configura­
flow-focused jets, however (see Table 1) requires precise current control tion, channels were also machined with the focusing air flow disabled at
in the milliampere range to deliver current densities suitable for EJM (10 210 A/cm2 and a feed rate of 0.35 mm/s. In flow focusing, the average
– 400 A/cm2) across different jet diameters, which is at the resolution jet speed is independent of flow rate and from Eq. (1) is given by vjet =
limit of the power supply used in this study (Keysight N5751A). Instead, √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
4Q/(πd2jet ) = 2Δpg /ρl ≈ 5.4 m/s at Δpg = 150 mbar. Hence, for the
the process was operated in controlled voltage mode. For every applied
conventional EJM the electrolyte flow rate was set to 10 ml/min to
flow rate, the theoretical jet diameter was calculated from Eq. (1) and
obtain a similar value (5.3 m/s for a nozzle internal diameter of
the expected resistance was estimated by assuming that the geometry of
200 µm).
the focused jet can be approximated by two cylinders with different
diameters, neglecting the contribution of electrode overpotentials.
Following the notation from Fig. 1 the estimated resistance Rjet is given
2.5. Surface metrology and analysis
by:
( )
4 T/2 + L T/2 + H Surface topography imaging of the machined samples was conducted
Rjet = + (2) using Alicona G4 focus variation microscope (FVM) with x20 magnifi­
πσ d2jet D21
cation objective (NA = 0.4). Analysis of the acquired topographic data
where σ is the electrolyte conductivity, djet is the diameter of the focused was performed using MountainsMap surface analysis software.
jet, D1 , T, H, L are the nozzle electrode inner diameter, throat thickness, The kerf widths reported in this work were averaged from 3 repeated
nozzle tip-to-orifice distance and stand-off distance respectively as per grooves at each condition in Table 1 with 5 measurements taken at
Fig. 1. The excitation voltage was chosen to ensure an approximately different positions along the groove length. The specific material volume
constant target current density of 210 A/cm2 across jet sizes estimated removed Vq per unit charge passed was also calculated from the cross-
sectional area of the machined grooves and the current measured from

4
I. Bisterov et al. Journal of Materials Processing Tech. 311 (2023) 117807

Fig. 3. Characterisation of the flow-focusing nozzle. (a) Optically measured jet diameters for the investigated parameter range and regions of stability. The
continuous contours are fitted to the experimental data (white markers), while the dashed ones are calculated from Eq. (1). (b) – (c) Optical images of the jet at the
annotated conditions.

the power supply during machining according to the relation Vq = Avf /I of the jet core and onset of break-up leading to a reduction of jet
where A is the cross-sectional area of the channel, vf is the applied feed cross-sectional area. This has previously enabled the measurement of
rate and I is the current. The standard uncertainty in the specific removal break-up length of liquid jets through electrical connectivity in studies
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
of airblast atomisers (Charalampous et al., 2016).
δV q was calculated from δV q = Vq (δI /I)2 + (δA /A)2 where δA is the
In the context of this work, the break-up length of the jet defines an
uncertainty in area taken as the standard deviation of the performed upper limit on the stand-off distance for electrochemical jet processing
measurements and δI is the standard uncertainty of the current mea­ since an intact jet is required. To quantify break-up length, least-squares
surement estimated to be 2.0 mA from the power supply specifications. fitting of the function y = Ax +B +CeDx was performed to each resistance
data set. The combination of linear and exponential function was chosen
3. Results and discussion to describe the two regions of the resistance response curve corre­
sponding to an intact jet and a jet undergoing break up respectively. The
3.1. The electrical resistance of flow-focused jets break-up location was taken as the point where the value of the second
derivative y ′ = CD2 eDx exceeded 300 kΩ/mm2. This value was arbi­

To study the electrical behaviour of the flow-focused jets, resistance trarily chosen here to define a certain degree of non-linearity and enable
response curves were acquired as a function of stand-off distance as quantitative comparison between resistance response curves at different
described in Section 2.2. The influence of flow rate Q and pressure dif­ conditions. The break-up locations identified through this method (an­
ferential Δpg across the orifice for the flow focusing configuration were notated with red dots in Fig. 2a,c) approached the focusing nozzle exit as
investigated and results were compared against optical images of the jet. the flow rate reduced demonstrating that a trade-off exists between
Response curves acquired at Δpg = 150 mbar and different flow rates are achievable jet diameter and useful working range. At 0.25 ml/min, a
shown in Fig. 2a. For flow rates from 0.5 ml/min to 2 ml/min, the curves linear resistance region could not be identified within the investigated
exhibited an approximately linear region below a certain stand-off dis­ range of 0.2–2 mm stand-off distance, and it was not possible to apply
tance whose slope increased with reduction in flow rate. This is indic­ the described fitting procedure. As seen in Fig. 2c, the jet at this con­
ative of jet contraction, since the reduction of the jet cross-sectional dition was unsteady along its full length. The average slope of the
leads to higher resistance per unit length (see Eq. (2)) However, upon response curves, obtained by linear fitting for stand-off distances smaller
increase of stand-off distance, the resistance of the jet increased sharply, than the break-up length for flow rates 0.5–2 ml/min and pressures 100
deviating from the initial linear behaviour. Observing optical images mbar – 300 mbar reduced with respect to jet diameter and demonstrated
obtained at the respective conditions (Fig. 2c), the generated flow- agreement with the expected trend for a cylindrical conductor (Fig. 2b).
focused jets exhibited lateral oscillations which increase in magnitude Electrical resistance measurement of the impinging jet proves to be a
downstream from the nozzle exit and ultimately lead to jet break-up. viable method to quantify the performance of the flow focusing tooling
This mode of flow focusing, known as “helical jetting” (Si et al., 2009) in-situ.
or “whipping”, when the whole jet is unstable (Blanco-Trejo et al.,
2020), has been previously observed to occur beyond a certain value of
applied pressure drop. Recently, Blanco-Trejo et al. (2020) demon­ 3.2. Operating regions of the flow-focusing nozzle
strated that the existence of this mode is strongly dependent on the
geometric parameters of the focusing nozzle design including the The operating parameter regions and achievable jet diameters of the
tip-to-orifice distance H and the shape of the throat, but general design flow focusing tool determined through optical imaging are summarised
rules have not been established yet. The acquired resistance curves in Fig. 3a. Jet diameters ranging from 164 µm ( ± 1 µm) at Q = 7.5 ml/
reflect the extent of oscillations observed optically, with the scatter min and Δpg = 100 mbar to 42 µm ( ± 4 µm) at Q = 0.5 ml/min and Δpg
evidently increasing at the more unstable conditions. The sharp increase = 150 mbar were observed over the investigated parameter window.
in resistance at larger stand-off distances corresponds to disintegration The latter value represents a 79% reduction of diameter compared to the

5
I. Bisterov et al. Journal of Materials Processing Tech. 311 (2023) 117807

Fig. 4. Resistance response curves acquired at pressure differentials Δpg of (a) 100 mbar, (b) 200 mbar, (c) 250 mbar, (d) 300 mbar.

Fig. 5. Break-up length of flow-focused jets. (a) Break-up length as a function of applied liquid flow rate and air pressure. Error bars represent standard deviation of
break-up measurements of 4 repeated resistance curves (b) Break-up length relative to jet diameter as a function of Weber number Weg.

6
I. Bisterov et al. Journal of Materials Processing Tech. 311 (2023) 117807

Fig. 6. Stand-off distance optimisation for machining. (a) Grooves machined at varying stand-off distance at 1 ml/min and 0.5 ml/min. (b)-(c) Influence of the work
piece surface on the air flow - the proximity of the surface affects the flow downstream of the orifice and disturbs the flow focusing effect.

200 µm ( ± 2 µm) jet generated without the flow focusing nozzle. The identifiable linear region, suggesting that high pressures promote
measured diameter values agreed with the expected values from Eq. (1) instability and earlier break-up.
(dashed contours in Fig. 3a) with relative error in the range of 0.3–13% The break-up length Lb quantified from the resistance curves as
and mean error of 5%. This demonstrates the capability enabled by the described in Section 3.1 decreased with reducing liquid flow rate and
developed tooling to vary the jet diameter to an arbitrary value within increasing pressure differential (see Fig. 5a). The break-up of flow-
the reported range by choosing the appropriate flow rate and pressure. focused jets and the more general problem of liquid jets with coaxially
In practice, however, the application of the smallest jet diameters to EJM flowing gas streams have been widely researched. A key dimensionless
may be challenging since stand-off distances of approximately 0.2 mm parameter is the aerodynamic Weber number (Lasheras and Hopfinger,
would be required between the nozzle and work piece to utilise the 2000) defined as
stable portion of the jet. The “whipping” region annotated in Fig. 3a
ρg v2g djet
represents the parameter domain where lateral oscillations were seen Weg = (3)
along the full length of the jet and the measurement of stable diameter γ
was not possible. Upon further reduction of flow rate to values lower
where ρg and vg are the density and velocity of the coaxial gas respec­
than 0.25 ml/min, an unstable behaviour consisting of periodic
detachment of liquid ligaments from the nozzle was observed, which is tively and γ is the surface tension of liquid-gas interface. The Weber
known as dripping and has been described in detail in fundamental number for the conducted experiments was calculated based on the
studies of flow focusing (Herrada et al., 2008). optically observed jet diameter and the pressure differential across the
Further resistance response curves were acquired at 100 – 300 mbar orifice considering that from Bernoulli’s law the average velocity of the
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
and 0.25–2 ml/min (Fig. 4) allow to characterise break-up length and gas vg = 2Δpg /ρg . The surface tension of the electrolyte-air interface
working range at the lower range of observed jet size. These revealed the was assumed to be 72 mN/m2. The break-up length was non-
same behaviour of linear increase of resistance at short stand-off dis­ dimensionalised by considering the ratio Lb/djet relative to jet diam­
tances followed by rapid increase upon jet disintegration. For Δpg = 100 eter. This exhibited an increasing trend with increasing aerodynamic
mbar and Q = 2 ml/min, the resistance curve appeared linear over the Weber number up to Weg ≈ 40 an reached a maximum value of
entire stand-off range, indicating a break-up length larger than 2 mm. approximately 20 (Fig. 5b). For Weg > 40 the break-up length decayed
Upon increase of Δpg , the high resistance regions appear to shift to lower again suggesting a change of break-up mode. The existence of transi­
stand-off values. At pressures above 200 mbar (see Fig. 4c-d), the tional Weber number is consistent with previous work by Gañán-Calvo
resistance curves at 0.5 ml/min became more unstable without an and Barrero (1999), but they identified a transtion from axisymmteric

7
I. Bisterov et al. Journal of Materials Processing Tech. 311 (2023) 117807

Fig. 7. Variable kerf through flow focusing. (a) Kerf width and specific material removal as a function of applied flow rate at a fixed pressure of 150 mbar. (b)
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
Exemplar average profiles extracted from the machined channels. (c) Specific volumetric material removal per unit charge as a function of vjet /djet . (d) Kerf width of
the machined channels normalised relative to jet diameter as function of jet diameter.

Rayleigh-type break-up driven by surface tension at low Weber number at high stand-off distances (larger than approximately 0.9 mm). This is
to a more irregular break-up pattern with non-axisymmetric helical jet likely caused by impaired electrical continuity due to jet break-up since
motion. As discussed earlier and shown in Fig. 2c., helical jet motions the break-up length was estimated to be 0.62 mm (S.D. = 0.05 mm). The
were observed this study even at low Weber numbers (12–39 for the jets overall shape of the groove appeared non-uniform, which could be the
in Fig. 2c) and this agrees with the observations of Si et al. (2009) over a result of the lateral oscillations of the jet. The observed widening of the
similar range for Δpg between 100 mbar and 200 mbar and Q between grooves at low stand-off distance can be explained by the interference of
0.3 ml/min and 1.7 ml/min. These differences in behaviour are likely the surface with the air flow from the focusing nozzle. As the jet of air
dependent on the geometry of the flow focusing devices employed. A impinges on the surface a stagnation region of high pressure and low
more detailed study of the geometric configuration could enable a more velocity is formed at the impingement location (Scholtz and Trass,
optimised design and a wider working range for the flow focusing 1970). When the tooling is sufficiently far from the surface (Fig. 6b), this
tooling in EJM, but is beyond the scope of this work. does not interfere with flow through the orifice and flow focusing can
occur. As the nozzle approaches the surface (Fig. 6c), the flow is blocked
by the surface and in the extreme scenario, the focusing orifice could lie
3.3. Variable tool size electrochemical jet machining
within the stagnation zone. It is expected that there will be a reduced
pressure drop across the orifice resulting in disturbance of the flow
A key difference between the proposed application to EJM and pre­
focusing effect and consequently larger jet sizes. This suggests that, in
vious research in flow focusing, is the fact that tooling in EJM is required
addition to the maximum stand-off defined by the break-up length of the
to function in close proximity to the work piece surface, which may have
jet, there is also minimum stand-off distance at which the flow focusing
an effect on the flow focusing effect and jet stability. Hence, the influ­
becomes significantly disturbed by the workpiece surface. The lower
ence of stand-off distance on machining at a fixed jet diameter was
limit can be controlled by reducing the orifice diameter relative to the
evaluated by machining channels with simultaneous movement of the Z-
stand-off distance.
axis as described in Section 2.4. The experiment was performed at a
Following the optimisation, stand-off distance was fixed at 0.9 mm as
pressure of 150 mbar and flow rates of 0.5 ml/min and 1 ml/min with a
an intermediate value shorter than the break-up length at the jet at 1 ml/
stand-off from 0.4 mm to 1.2 mm, to cover the working range defined
min, but sufficiently large so that the focusing effect is not disturbed by
through resistance measurements (average break-up length for 1 ml/
the work piece surface (Fig. 6a). Machining of straight channels was
min is 1.24 mm with S.D. = 0.02 mm, see Fig. 2c). For both investigated
undertaken using the parameters listed in Table 1. The average kerf
flow rates, the machined groove visibly became wider and more un­
width reduced from 308 µm (S.D. = 5 µm) to 154 µm (S.D. = 7 µm) upon
stable when the nozzle was positioned closer than approximately
reduction of flow rate from 7.5 to 1 ml/min (Fig. 7a-b). The channel
0.7 mm (Fig. 6a). For the 0.5 ml/min no material removal was detected

8
I. Bisterov et al. Journal of Materials Processing Tech. 311 (2023) 117807

Fig. 8. “The Starry Night” engraved on a stainless-steel work piece through flow-focused EJM. (a) Optical image of the sample. (b) 3-diemnsional surface topography
reconstruction.

machined without flow focusing had a kerf width of 332 µm (S.D. = variation of jet size through the presented method would always result
6 µm), demonstrating 54% reduction through the developed tooling. in controlled variation of machining resolution.
With reduction of jet diameter, the volumetric material removal rate is One constraint that the flow focusing approach places on EJM is that
significantly reduced. However, it is interesting that the specific volu­ the jet speed cannot be changed independently of jet diameter, since
metric material removal per unit charge also reduced (Fig. 7a) along they are both defined by the applied pressure differential to the gas with
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
with a reduction in the depth of the machined channels (see Fig. 7b) vjet ≈ 2Δpg /ρl . The jet speeds reported in conventional EJM are often
demonstrating a decrease in dissolution efficiency. The specific removal higher at 10 – 20 m/s than the ones applied in the conducted machining
reduced from 16.7 × 10− 3 mm3/C (standard uncertainty δ = 0.6 ×10− 3 experiments (Speidel et al., 2022). However, jet speeds in excess of
mm3/C) for a flow rate of 7.5 ml/min to 10.6 × 10− 3 mm3/C 70 m/s (with applied pressures Δpg > 30 bar) however have been re­
(δ = 2.1 ×10− 3 mm3/C) for a flow rate of 1 ml/min. ported in flow focusing for serial femtosecond crystallography applica­
The specific removal measured from the conventional EJM config­ tions (Grünbein et al., 2018). This suggests that optimisation of the
uration (disabled focusing air flow, measured current density J ≈ 216 A/ geometry of the flow focusing tooling may be possible for an extended
cm2, δ = 7 A/cm2) at 10 ml/min and equivalent stand-off distance was operating pressure range to deliver stable jets of sufficient length at the
17.5 × 10− 3 mm3/C (δ = 0.5 ×10− 3 mm3/C). A possible explanation for desired jet speed for EJM. For the specific material-electrolyte system
this reduction in dissolution efficiency is that the transport-controlled used in this study, this not likely to be beneficial since Fig. 7c suggests
oxygen evolution reaction (OER), which is known to occur in parallel that this would further reduce dissolution efficiency due to a more
with the anodic dissolution of iron-based alloys in nitrate solutions, is increased rate of the OER. An increase of jet speed at a fixed current
proceeding at an increased rate (consuming a greater proportion of the density however leads to lower temperature increase in the IEG, so this
current). As discussed by Lohrengel et al. (2016), the OER is understood may have additional effects on the anodic reaction rates.
to be limited by the depletion of water (dehydration) at the anode sur­ Through flow focusing, the ability to vary machining resolution for a
face when electrochemical machining products are formed. The rate of specific material is made available to the EJM practitioner through
transport-controlled processes occurring at the electrode under an simple modification of electrolyte flow rate and air pressure. This re­
impinging jet has been shown by Chin and Tsang (1978) to be propor­ duces the need for different tool nozzles and has the potential to
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
tional to vjet /djet , where vjet is the average jet velocity. The measured significantly accelerate machining operations where features of
specific material removal is plotted against this quantity in Fig. 7c, different scale are required. Flow rate and pressure can be automatically
where the jet speed was calculated from vjet = 4Q/(π d2jet ) and was in the controlled by the machine tool control programme as a function of tool
range of 5.0 – 6.7 m/s. The observed decreasing trend could indicate path, hence eliminating process interruptions for tool changes. To
that there is an increased supply of water molecules to the electrode showcase this new capability, a replica of the painting “The Starry
surface and respectively higher limiting current density for the OER, Night” is engraved in a 16 mm × 13 mm area on a stainless-steel
leading to reduced dissolution efficiency of the work piece material workpiece (Fig. 8). Different sections of the sample were machined at
when the flow-focused jet diameter decreased. Decreasing jet diameter differing electrolyte flow rates at a fixed air pressure of 150 mbar and the
also resulted in an increase of non-dimensional kerf widths relative to jet respective parameters from Table 1 representing different “brushes”
diameter, which varied from 1.88 (S.D. = 0.03) to 2.52 (S.D. = 0.11) (see required dependent on the feature size. More advanced pump control
Fig. 7d), demonstrating that kerf width does not scale linearly with jet could enable the dynamic variation of flow rate during machining
diameter and could also be the result of the discrete electrochemical allowing the fabrication of channels with a controlled non-uniform
processes occurring at the anode surface. Such effects are specific to the cross-section, which will significantly expand the surface texturing ca­
chosen work piece material – electrolyte combination, which is known pabilities of EJM.
to promote passive layer formation and oxygen evolution at the anode,
so the extent of kerf reductions achieved through flow focusing would 4. Conclusions
vary for different surface chemistries. However, in principle the
A novel variable-size tool for electrochemical jet machining based on

9
I. Bisterov et al. Journal of Materials Processing Tech. 311 (2023) 117807

the flow focusing principle was proposed. In this work, its potential to Blanco-Trejo, S., Herrada, M.A., Gañán-Calvo, A.M., Rubio, A., Cabezas, M.G.,
Montanero, J.M., 2020. Whipping in gaseous flow focusing. Int. J. Multiph. Flow.
increase the precision and flexibility of the process was demonstrated.
130, 103367.
Compared to the conventional tool nozzle, a 54% reduction in machined Chapman, H.N., Fromme, P., Barty, A., White, T.A., Kirian, R.A., Aquila, A., Hunter, M.S.,
kerf width was observed at a liquid flow rate of 1 ml/min and pressure Schulz, J., DePonte, D.P., Weierstall, U., Doak, R.B., Maia, F.R.N.C., Martin, A.V.,
drop of 150 mbar enabling the variation of tool size through selection of Schlichting, I., Lomb, L., Coppola, N., Shoeman, R.L., Epp, S.W., Hartmann, R.,
Rolles, D., Rudenko, A., Foucar, L., Kimmel, N., Weidenspointner, G., Holl, P.,
electrolyte flow rate or pressure of the focusing gas independent of the Liang, M., Barthelmess, M., Caleman, C., Boutet, S., Bogan, M.J., Krzywinski, J.,
nozzle electrode diameter. Electrical resistance measurements were Bostedt, C., Bajt, S., Gumprecht, L., Rudek, B., Erk, B., Schmidt, C., Hömke, A.,
shown to be a viable method for characterising the tooling performance Reich, C., Pietschner, D., Strüder, L., Hauser, G., Gorke, H., Ullrich, J., Herrmann, S.,
Schaller, G., Schopper, F., Soltau, H., Kühnel, K.-U., Messerschmidt, M., Bozek, J.D.,
in-situ and quantifying jet diameter and break-up length. The observed Hau-Riege, S.P., Frank, M., Hampton, C.Y., Sierra, R.G., Starodub, D., Williams, G.J.,
stable jet lengths were in the range of 13–20 jet diameters resulting in a Hajdu, J., Timneanu, N., Seibert, M.M., Andreasson, J., Rocker, A., Jönsson, O.,
trade-off between achievable jet contraction and working range of the Svenda, M., Stern, S., Nass, K., Andritschke, R., Schröter, C.-D., Krasniqi, F., Bott, M.,
Schmidt, K.E., Wang, X., Grotjohann, I., Holton, J.M., Barends, T.R.M., Neutze, R.,
tooling. The smallest jet size successfully applied for machining was Marchesini, S., Fromme, R., Schorb, S., Rupp, D., Adolph, M., Gorkhover, T.,
61 µm ( ± 2 µm). Andersson, I., Hirsemann, H., Potdevin, G., Graafsma, H., Nilsson, B., Spence, J.C.H.,
Further research would involve the optimisation of the tooling 2011. Femtosecond X-ray protein nanocrystallography. Nature 470, 73–77.
Charalampous, G., Hadjiyiannis, C., Hardalupas, Y., 2016. Comparative measurement of
design, including the application of flow conditioning to reduce distur­ the breakup length of liquid jets in airblast atomisers using optical connectivity,
bances and investigations to establish the influence of geometric pa­ electrical connectivity and shadowgraphy. Measurement 89, 288–299.
rameters such as orifice diameter, nozzle tip-to-orifice distance and Chin, D.T., Tsang, C.H., 1978. Mass transfer to an impinging jet electrode.
J. Electrochem. Soc. 125, 1461-1461.
overall focusing throat shape on the fluid dynamic performance and
Clare, A.T., Speidel, A., Bisterov, I., Jackson-Crisp, A., Mitchell-Smith, J., 2018. Precision
stability. This could further expand the precision, working range and enhanced electrochemical jet processing. CIRP Ann. 67, 205–208.
applicability of the proposed concept. The resulting surface texture after Datta, M., Romankiw, L.T., Vigliotti, D.R., von Gutfeld, R.J., 1989. Jet and laser-jet
machining and quality of the cut as jet size is further reduced relative to electrochemical micromachining of nickel and steel. J. Electrochem. Soc. 136,
2251–2256.
the grain size of the work piece material are also an interesting research Deponte, D.P., McKeown, J.T., Weierstall, U., Doak, R.B., Spence, J.C.H., 2011. Towards
avenue. ETEM serial crystallography: electron diffraction from liquid jets. Ultramicroscopy
111, 824–827.
Gañán-Calvo, A.M., 1998. Generation of steady liquid microthreads and micron-sized
CRediT authorship contribution statement monodisperse sprays in gas streams. Phys. Rev. Lett. 80, 285–288.
Gañán-Calvo, A.M., Barrero, A., 1999. A novel pneumatic technique to generate steady
Ivan Bisterov: Conceptualization, Investigation, Methodology, capillary microjets. J. Aerosol Sci. 30, 117–125.
Gañán-Calvo, A.M., González-Prieto, R., Riesco-Chueca, P., Herrada, M.A., Flores-
Writing – original draft. Sidahmed Abayzeed: Methodology, Writing – Mosquera, M., 2007. Focusing capillary jets close to the continuum limit. Nat. Phys.
review & editing. Alistair Speidel: Supervision, Writing – review & 3, 737–742.
editing. Mirco Magnini: Writing – review & editing. Mohamed Grünbein, M.L., Bielecki, J., Gorel, A., Stricker, M., Bean, R., Cammarata, M., Dörner, K.,
Fröhlich, L., Hartmann, E., Hauf, S., Hilpert, M., Kim, Y., Kloos, M., Letrun, R.,
Zubayr: Writing – review & editing. Adam T. Clare: Conceptualization, Messerschmidt, M., Mills, G., Nass Kovacs, G., Ramilli, M., Roome, C.M., Sato, T.,
Supervision, Writing – review & editing. Scholz, M., Sliwa, M., Sztuk-Dambietz, J., Weik, M., Weinhausen, B., Al-Qudami, N.,
Boukhelef, D., Brockhauser, S., Ehsan, W., Emons, M., Esenov, S., Fangohr, H.,
Kaukher, A., Kluyver, T., Lederer, M., Maia, L., Manetti, M., Michelat, T.,
Declaration of Competing Interest Münnich, A., Pallas, F., Palmer, G., Previtali, G., Raab, N., Silenzi, A., Szuba, J.,
Venkatesan, S., Wrona, K., Zhu, J., Doak, R.B., Shoeman, R.L., Foucar, L.,
Colletier, J.-P., Mancuso, A.P., Barends, T.R.M., Stan, C.A., Schlichting, I., 2018.
The authors declare that they have no known competing financial
Megahertz data collection from protein microcrystals at an X-ray free-electron laser.
interests or personal relationships that could have appeared to influence Nat. Commun. 9, 3487.
the work reported in this paper. Herrada, M.A., Gañán-Calvo, A.M., Ojeda-Monge, A., Bluth, B., Riesco-Chueca, P., 2008.
Liquid flow focused by a gas: Jetting, dripping, and recirculation. Phys. Rev. E 78,
036323.
Data Availability Keohane, K., Brennan, D., Galvin, P., Griffin, B.T., 2014. Silicon microfluidic flow
focusing devices for the production of size-controlled PLGA based drug loaded
Data will be made available on request. microparticles. Int. J. Pharm. 467, 60–69.
Lasheras, J.C., Hopfinger, E.J., 2000. Liquid jet instability and atomization in a coaxial
gas stream. Annu. Rev. Fluid Mech. 32, 275–308.
Acknowledgements Li, T., Yan, X., Fang, X., Jin, P., Li, J., Rabbi, K.F., Miljkovic, N., 2021. In situ jet
electrolyte micromachining and additive manufacturing. Appl. Phys. Lett. 119,
171602.
The authors would like to acknowledge Alexander Jackson-Crisp of Lohrengel, M.M., Rataj, K.P., Münninghoff, T., 2016. Electrochemical
the Advanced Component Engineering Laboratory at the University of machining—mechanisms of anodic dissolution. Electrochim. Acta 201, 348–353.
Nottingham for his input to the design of the reported tooling and its Lutey, A.H.A., Jing, H., Romoli, L., Kunieda, M., 2021. Electrolyte Jet Machining (EJM)
of antibacterial surfaces. Precis. Eng. 70, 145–154.
manufacture, along with the general technical support of this work. This Mitchell-Smith, J., Speidel, A., Bisterov, I., Clare, A.T., 2018. Electrolyte multiplexing in
work was supported by the Engineering and Physical Sciences Research electrochemical jet processing. Procedia CIRP 68, 483–487.
Council (UK) [grant numbers: EP/W018942/1 and EP/L016206/1] Nelson, G., Kirian, R.A., Weierstall, U., Zatsepin, N.A., Faragó, T., Baumbach, T.,
Wilde, F., Niesler, F.B.P., Zimmer, B., Ishigami, I., Hikita, M., Bajt, S., Yeh, S.-R.,
through Sustainable Manufacturing and the Centre for Doctoral Training Rousseau, D.L., Chapman, H.N., Spence, J.C.H., Heymann, M., 2016. Three-
in Innovative Materials Processing. Professor Clare would like to dimensional-printed gas dynamic virtual nozzles for x-ray laser sample delivery. Opt.
acknowledge the kind support of the Royal Academy of Engineering, Express 24, 11515–11530.
Ponce-Torres, A., Vega, E.J., Castrejón-Pita, A.A., Montanero, J.M., 2017. Smooth
RCSRF1920927. printing of viscoelastic microfilms with a flow focusing ejector. J. Non-Newton. Fluid
Mech. 249, 1–7.
References Ponce-Torres, A., Ortega, E., Rubio, M., Rubio, A., Vega, E.J., Montanero, J.M., 2019.
Gaseous flow focusing for spinning micro and nanofibers. Polymer 178, 121623.
Saxena, K.K., Qian, J., Reynaerts, D., 2018. A review on process capabilities of
Acero, A.J., Ferrera, C., Montanero, J.M., Gañán-Calvo, A.M., 2012. Focusing liquid
electrochemical micromachining and its hybrid variants. Int. J. Mach. Tools Manuf.
microjets with nozzles. J. Micromech. Microeng. 22, 065011.
127, 28–56.
Acero, A.J., Rebollo-Muñoz, N., Montanero, J.M., Gañán-Calvo, A.M., Vega, E.J., 2013.
Saxena, K.K., Qian, J., Reynaerts, D., 2020. A tool-based hybrid laser-electrochemical
A new flow focusing technique to produce very thin jets. J. Micromech. Microeng.
micromachining process: Experimental investigations and synergistic effects. Int. J.
23, 065009.
Mach. Tools Manuf. 155, 103569-103569.
Beyerlein, K.R., Adriano, L., Heymann, M., Kirian, R., Knoška, J., Wilde, F., Chapman, H.
Scholtz, M.T., Trass, O., 1970. Mass transfer in a nonuniform impinging jet: Part I.
N., Bajt, S., 2015. Ceramic micro-injection molded nozzles for serial femtosecond
Stagnation flow-velocity and pressure distribution. AIChE J. 16, 82–90.
crystallography sample delivery. Rev. Sci. Instrum. 86, 125104.
Si, T., Li, F., Yin, X.-Y., Yin, X.-Z., 2009. Modes in flow focusing and instability of coaxial
Bisterov, I., Abayzeed, S., Speidel, A., Clare, A.T., 2022. On-machine measurement with
liquid–gas jets. J. Fluid Mech. 629, 1–23.
an electrochemical jet machine tool. Int. J. Mach. Tools Manuf. 174, 103859.

10
I. Bisterov et al. Journal of Materials Processing Tech. 311 (2023) 117807

Speidel, A., Bisterov, I., Saxena, K.K., Zubayr, M., Reynaerts, D., Natsu, W., Clare, A.T., Wu, M., Liu, J., He, J., Chen, X., Guo, Z., 2020. Fabrication of surface microstructures by
2022. Electrochemical jet manufacturing technology: From fundamentals to mask electrolyte jet machining. Int. J. Mach. Tools Manuf. 148, 103471.
application. Int. J. Mach. Tools Manuf. 180, 103931. Xinchao, L., Pingmei, M., Xinmin, Z., Wei, W., Yanhua, Z., Ge, Q., Xingshuai, Z., Shen, N.,
Vega, E.J., Montanero, J.M., Herrada, M.A., Gañán-Calvo, A.M., 2010. Global and local 2020. Compressed air-film encircling jet electrodeposition with high deposition
instability of flow focusing: The influence of the geometry. Phys. Fluids 22, 064105. accuracy. J. Electrochem. Soc. 167, 102502.
Walker, J.C., Cinti, S., Kamps, T.J., Mitchell-Smith, J., Clare, A.T., 2019. Influence of Yang, X., Liu, X., Lu, Y., Zhou, S., Gao, M., Song, J., Xu, W., 2016. Controlling the
contact area on the sliding friction and wear behaviour of an electrochemical jet adhesion of superhydrophobic surfaces using electrolyte jet machining techniques.
textured Al-Si alloy. Wear 426–427, 1336–1344. Sci. Rep. 6, 23985.
Wei, W., Pingmei, M., Xinmin, Z., Xinchao, L., Yunyan, Z., Shen, N., Sansan, A., 2022.
Additive manufacturing of three-dimensional intricate microfeatures by electrolyte-
column localized electrochemical deposition. Addit. Manuf. 50, 102582.

11

You might also like