Download as pdf or txt
Download as pdf or txt
You are on page 1of 33

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/326299317

Recent advances and challenges of abrasive jet machining

Article  in  CIRP Journal of Manufacturing Science and Technology · July 2018


DOI: 10.1016/j.cirpj.2018.06.001

CITATIONS READS

66 5,543

2 authors:

Ruslan Melentiev Fengzhou Fang


King Abdullah University of Science and Technology University College Dublin
16 PUBLICATIONS   221 CITATIONS    398 PUBLICATIONS   6,414 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Atomic Scale Manufacturing: A first hand understanding View project

AR diffractive waveguide View project

All content following this page was uploaded by Ruslan Melentiev on 07 February 2021.

The user has requested enhancement of the downloaded file.


Review

Recent advances and challenges of abrasive jet machining


Ruslan Melentiev1, Fengzhou Fang 1, 2*

1. Centre of Micro/Nano Manufacturing Technology, University College Dublin, Dublin, Ireland;


2. Centre of Micro/Nano Manufacturing Technology, Tianjin University, Tianjin, China;

* Corresponding email: fengzhou.fang@ucd.ie

Abstract
Abrasive jet machining (AJM) is a manufacturing technology based on erosion localization and intensification.
AJM has a progressively important influence on the machining technology market. Over the past 20 years, there
has been an exponential growth in the number of papers that discuss AJM. Various innovations and process
developments such as intermittent, submerged, thermally assisted and other jet conditions were proposed. This
paper examines AJM’s technological advantages and the variety of machining operations in different industries
where AJM is utilized. Particular attention is devoted to the micro-texturing capabilities of powder blasting and
its application in tribology. New evidence of ductile and brittle material removal mechanisms are reviewed
together with recently discovered elastic removal mode. The effects of hydraulic, abrasive and machining
parameters on particles kinetic energy, machined surface roughness and footprint size are described in detail.
Nozzle wear has a strong dependence on nozzle materials, its geometry, particles size, hardness, and flow rate.
The trend of AJM development is a shift from macro to micro scale. Improvements in micro-machining resolution,
process controlling and erosion prediction are current challenges facing AJM.

Keywords: Abrasive jet machining, powder blasting, micromachining, material removal mechanism, process
parameters, nozzle wear.

Contents

1. Introduction.................................................................................................................................................... 2
2. Approaches of abrasive jet machining ........................................................................................................... 3
2.1. Abrasive Air Jet Machining .................................................................................................................. 3
2.2. Abrasive Water Jet Machining .............................................................................................................. 3
2.3. Magnetorheological Jet Polishing ......................................................................................................... 4
2.4. Cryogenic Assisted Abrasive Jet Machining ......................................................................................... 5
2.5. Thermally Enhanced Abrasive Jet Machining ...................................................................................... 5
2.6. Cavitation Assisted Fluid Jet Polishing ................................................................................................. 6
2.7. Ice Assisted Jet Machining .................................................................................................................... 6
2.8. Intermittent Jet Conditions .................................................................................................................... 6
2.9. Submerged Jet Conditions..................................................................................................................... 6
2.10. Multi Energy Jet Conditions ............................................................................................................. 7
3. Capabilities and application ........................................................................................................................... 7
3.1. Technological advantages and drawbacks ............................................................................................ 7
3.2. Assignment and areas of application ..................................................................................................... 8
3.3. Texturing capabilities ............................................................................................................................ 9
3.4. Outlook for tribology .......................................................................................................................... 10
4. Material removal mechanism ....................................................................................................................... 11
4.1. Brittle Material Removal Mode .......................................................................................................... 11
4.2. Plastic Material Removal Mode .......................................................................................................... 12
4.3. Elastic Removal Mode ........................................................................................................................ 13
5. Influence of process parameters ................................................................................................................... 14
5.1. Hydraulic parameters .......................................................................................................................... 14
5.1.1. Nozzle pressure .......................................................................................................................... 14
5.1.2. Jet velocity.................................................................................................................................. 15
5.2. Machining parameters ......................................................................................................................... 16
5.2.1. Traverse speed ............................................................................................................................ 16
5.2.2. Feed step ..................................................................................................................................... 16
5.2.3. Stand-off-distance ...................................................................................................................... 16
5.2.4. Incidence angle ........................................................................................................................... 17
5.3. Abrasive parameters ............................................................................................................................ 17
5.3.1. Abrasive size .............................................................................................................................. 17
5.3.2. Abrasive hardness....................................................................................................................... 18
5.3.3. Flow rate ..................................................................................................................................... 19
5.4. Process parameters summary .............................................................................................................. 19
6. Nozzle wear issue ........................................................................................................................................ 20
6.1. Wear mechanism ................................................................................................................................. 20
6.2. Influence of nozzle material ................................................................................................................ 21
6.3. Influence of nozzle geometry .............................................................................................................. 22
6.4. Influence of abrasives and process parameters ................................................................................... 22
7. Conclusions and prospect ............................................................................................................................ 23
8. Acknowledgement ....................................................................................................................................... 24
References ............................................................................................................................................................ 24

1. Introduction

In the 1930s, a low-pressure water jet system was patented and successfully used to cut paper [1]. Twenty
years later, a high-pressure hydraulic seal from aviation industry was adopted to water jet machining, that
noticeably increased the process productivity [2]. The continuous increase of working pressure in the next few
decades allowed the cutting of hard alloys and carbides. On the other hand, a high pressure led to severe nozzle
wear, making abrasive jet machining (AJM) economically non-competitive. From the 1970s, after ceramic nozzles
were introduced, abrasive jet systems became commercially available and, within a short span of time, became
the industrial mainstream and were mainly utilised for cutting and cleaning purposes. Further developments of
AJM technology have been made, mainly based on material science progress and CNC conception. In the 21st
century, AJM development deviated its track to technology miniaturization, wherein the nozzle diameter plunged
from macro to micro scale. Today, sapphire orifice, super-hard abrasives and reliable high-pressure pumps
combined with a 6-axis, precisely manage and process monitored systems, making AJM one of the most promising
micro-manufacturing technologies despite the fact that it has been used for a century. In the last 20 years, there is
a solid growing trend of industrial interest in micro-AJM. The obvious reflection of industrial demands can be
seen in a commutative volume of research activity in the area. Since 2000, there has been an exponential growth
in the number of publications displayed by Engineering village and Science Direct databases on the request:
“abrasive jet machining”.
Previous articles [3–7] have described the recent technological state of AJM. Nevertheless, the review
of Chen et al. [3] focused on polishing capabilities of AJM. Verma et al. [4] and Syazwani et al. [5] reviewed the
nozzle wear in abrasive waterjet machining (AWJM) separately. Molitoris et al. [6] reported on developments of
abrasive water suspension jets. Kalpana et al. [7] analyzed only the process monitoring methods. Taking this
research into account, it is necessary to fill the gaps and provide a comprehensive review on the state of the art of
AJM, including its technological strengths and weaknesses, analysis of AJM developments and material removal
mechanism, the influence of process parameters on surface integrity, texturing capabilities and nozzle wear in
abrasive air jet machining (AAJM).
The aim of this work is to fill the gaps in previous articles, by highlighting the main aspects of the
technology and representing the most relevant and latest findings among experimental and theoretical
investigations. Firstly, two chapters cover technological advantages, industrial applications, and diversity of AJM
approach. Chapters 3 - 5 give a review of material removal mechanisms, process parameters influence and nozzle
wear focusing on AAJM. The structure of the paper forms a general view of AJM’s current technological state
and detailed assessment of AAJM. Technology problems, their potential solutions, and future prospects are
discussed in the conclusion. This review does not cover the modelling of abrasive jet processes. The authors
believe that the progress in particle velocity modelling, prediction of material removal rate and surface evolution
are worthy of a separate review.
2. Approaches of abrasive jet machining

The variety of industrial demands for manufacturing of different parts with a specific geometry, surface
roughness, and integrity led to several AJM modifications. Apart from well-known abrasive air and waterjet
methods, a magnetorheological jet machining (MJM) was invented for superfinishing of precision optics. To
change material removal rate (MRR), an abrasive jet can be assisted by cryogenic or high temperatures, air
cavitation, etc. For other purposes, such as deep grooving, noise and pollution reduction or large area machining,
corresponding proposals were made on the intermittent and submerged jet conditions or energy jet division. This
chapter briefly introduces the AJM approaches, its benefits, and latest advances.

2.1. Abrasive Air Jet Machining

Abrasive air jet machining (AAJM) and micro AAJM are generally categorized as blasting [8]. While
conventional shot blasting is applied mainly to surface cold-hardening, AAJM is aimed at material removal. When
the nozzle exit diameter and machining feature is less than 1 mm, the process can be called micro-AAJM, micro-
blasting or powder blasting. A typical AAJM system is shown in Figure 1Error! Reference source not found..
Compressed air is cleaned and dried through the in-system filter and desiccator. Then, air moves through the
pressure adjusting valve to the blasting gun where it is mixed with abrasive particles and is blasted at the workpiece
with controlled pressure. Worked out particles and chips are eliminated from the machining area by the dust
catcher. By impacting the substrate, particles produce small fractures or plastic extrusion, depending on material
properties and process parameters. After impact, gas flow carries away both the abrasives and chips. Micro-AAJM
is mainly devoted to the fabrication of micro holes and surface patterns on brittle substrates such as silicon,
ceramics, metallic carbides and glasses. Processing of polymers and non-ferrous metals are often accompanied
with negative effects of abrasives embedding into the machined surface (detailed in section “Abrasive size”).
Some successful attempts to machine polymers using fine and fast particles were made with cryogenic assistance
(see the corresponding section).

4 3 2 1

6
9

8
7 Legend:
1. Air compressor
2. Air filter
3. Air desiccator
4. Pressure valve
5. Abrasive feeder
6. Nozzle
7. Workpiece
8. Dust absorber
9. Blasting chamber

Figure 1. Schema of AAJM system

2.2. Abrasive Water Jet Machining

Abrasive waterjet machining (AWJM), as a technological method, was developed in 1930 in the mining
industry [6]. Compared with AWJM, the abrasive slurry jet machining (ASJM) or sometimes abrasive water
suspension jet (AWSJM), operates at roughly two orders of magnitude lower pressure and uses a premixed slurry
concentration. The method is based on the same principle as AAJM, with a difference in fluid viscosity. Jet
characteristics are improved by water’s one hundred times greater viscosity than air. Water combined with
abrasive particles forms slurry. The slurry is pressurized and delivered through a nozzle to the machined surface.
Water jet machining provides a lot of advantages over AAJM, including: Nano-scale surface roughness [9]; more
distant and accurate footprint [10]; increased cooling of the workpiece and efficient recovery of slurry [11,12].
However, impact velocities of the particles in AAJM and AWJM are not equal because of the strong decelerating
effect in the water stagnation zone and the strong influence of the fluid viscosity on particle trajectory. The
stagnation zone is the area of significant deceleration of abrasive particles when approaching the surface. The
cross-section of numerically simulated water jet velocity is shown in Figure 2. Crossing the stagnation zone, a
particle decelerates and loses its kinetic energy. Under comparable process parameters, AAJM provides higher
MRR due to its less expressed stagnation zone [13]. Matsumura et al. [14] proposed the controlling of the
stagnation area using tapered V-shaped masks. Sharper taper angle provides larger impingement angles that allow
particles to collide onto the surface almost vertically. Kowsari et al. [15] showed that improvements in the control
of the eroding zone can be achieved using highly viscous soybean oil as an energy carrier fluid to prevent the
secondary particle impact in the primary footprint. Numerous works were devoted to process parameters
optimization [16–19], numerical simulations of particles impact and erosion [56, 57], surface topology and
roughness prediction [21–23]. The disadvantages of the AWJM is the presence of fluid, which produces a
deleterious effect on crack propagation, particularly in ceramics, [24], as well as a hydrodynamic lubrication effect
[25].

Figure 2. Flow velocity distribution [14].

2.3. Magnetorheological Jet Polishing

Inherent water jet instability results in an unstable MRR and inconstant size of machining spot that limits
the application of AWJM for precision surface finishing. The aerodynamic disturbance depends directly on the
Reynolds number, which is directly proportional to jet velocity and diameter, but inversely proportional to jet
viscosity. The increment of jet viscosity leads to higher pumping power of the system and makes this method
impractical for high-speed jetting. To increase the jet velocity without a loss in its stability, a new
magnetorheological finishing (MRF) method was proposed [26–28]. MRF method is based on
magnetorheological fluid, which is a water mixture of abrasives and dispersed ferromagnetic particles with
chemical additives. Under a magnetic field, ferromagnetic suspension changes its mechanical properties literally
to the plastic pattern, depending on characteristics of the magnetic field applied. Under magnetization, particles
are organized into a spatial structure, which have a higher yield stress among other things. MR liquid inside the
nozzle is suppressed and internal jet vacancies are filled. As a result, a highly chained and collimated fluid jet can
be obtained inside and beyond the nozzle[29]. Despite magnetic field weakening, the suppressed and continually
filled state of the jet is preserved as it travels further from the nozzle exit [30]. Therefore, an initial magnetic jet
stabilization at the nozzle may constrain its coherence for further prolonged distance with irrelevant disturbance.
The comparison of water and MR jet stability is shown in Figure 3. In contrast with water, where jet diameter
remains constant for a few nozzle diameters, MR liquid is more viscous and so remains stable for up to 10
diameters. At the same time, the MR liquid within an activated magnetic field demonstrates persistent jet circle
for 200 diameters, which can be up to 0.5 meters depending on the jet diameter. Experimentation with stand-off
distances (SOD), 50 mm and 150 mm, shows the maximum variation of material removal for only 6.5% [28].
That makes MRF a particularly interesting technology for machining of hard approachable surfaces. It is also
established [31–33] that by using MRF, a high precision surface with roughness of less than 1 nm rms is achievable
for a number of brittle materials like single crystals, ceramics, and glasses. The drawback of this technology is the
technical complexity and costliness.
Figure 3. Jet stability snapshot (jet velocity = 30 m/s, nozzle diameter 2 mm) [28].

2.4. Cryogenic Assisted Abrasive Jet Machining

Relatively soft metals and alloys cannot be precisely machined with a classical AAJM because of definite
surface integrity casualties. Small particles with a high velocity and large kinetic energy penetrate the substrate,
especially when particles impact the surface at a normal angle. Numerous materials cannot be chipped in a brittle
way due to their low hardness. In such situations, cryogenic assistance to the machining process might
substantially change the material removal mechanism. Yuvaraj and Kumar [34] compared the surface integrity of
AA5083-H32 aluminum alloy after AWJ and CAAJM. The liquid nitrogen at −196 °C was delivered to the impact
zone with compressed air. An increase in substrate hardness through the deep freezing led to a drastic decrease in
the number of abrasives embedded in the machined surface. A more repeatable roughness profile and smaller
peaks and valleys differences were achieved as well. In addition, the strain hardening effect due to grain
refinement by liquid nitrogen almost doubled the microhardness of CAAJ machined surface with all tested process
parameters. Getu et al. [35] cooled the polydimethylsiloxane (PDMS) surface by liquid nitrogen and machined
with AAJM simultaneously. At −150 °C, the maximum erosion was achieved at a normal impact angle that implies
a brittle removal mode, since in a ductile mode a maximum would arise within a range of 15°- 30°. The approach
to machining soft materials by brittle fracturing was discovered. In subsequent studies [36,37], authors used a heat
exchanger with liquid nitrogen to cool the air jet instead of a workpiece. A tube formed of materials with high and
low thermal conductivity were applied inside and outside of LN2 heat exchanger, correspondingly. With the jet
temperature of −80 °C, the abrasives were embedded abundantly into the polymer surface, albeit particles were
almost absent at −180 °C. Erosion mechanism was the same during processing at – 80°C and –127°C, but changes
noticeably at –180°C from the ductile to brittle mode. The relation of volume removal per unit of kinetic energy
of incident particles (mm3/J) significantly increased with the maximum between 30° and 60° of impact angle. This
implies that, even at −180 °C, there is a combination of brittle and plastic removal mechanisms. Thus, deeper
cooling can be considered to obtain a clean brittle removal mechanism. The first drawback of the proposed method
is a multiple increase of energy consumption by the cooling system with liquid nitrogen. The second is a heavy
microcracking of elastomer surface after cooling to −180 °C in virtue of thermal strain. However, Gradeen et al.
[36] indicated that the cold cracks might weaken the surface for further crack propagation, which contributes to
consequent erosion.

2.5. Thermally Enhanced Abrasive Jet Machining

Thermally enhanced AWJM was explored by Patel and Tandon [38]. Before blasting, the workpiece was
heated up to 850 ºC by the oxy-acetylene gas-welding torch, for MRR acceleration of hard-to-machine materials.
The high temperature in the cutting zone reduces the yield strength and hardness of the workpiece, which causes
its softening and a consequent reduction in the amount of cutting energy required. Statistical results depicted the
increase of MRR machining Inconel 718 and Titanium (Ti6Al4V). Substrate properties were almost unimpaired.
Even when the material undergoes thermo-mechanical rapid cooling, no residual stresses and cracks were
observed. However, some micro-cracks on the machined surface were found when mild steel (MS-A36) was
machined. Patel and Tandon remarked on difficulties in spatial temperature maintenance using welding setup and
recommend considering other approaches to local workpiece heating.

2.6. Cavitation Assisted Fluid Jet Polishing

RMR of micro-AWJM is limited to abrasive size and process pressure [39,40] increase, which leads to
continuous degradation of surface integrity [41]. Yu et al [42] proposed so-called air-driven fluid jet machining,
where the slurry is dropped out of the tank due to the Venturi effect in the atomizer, and then mixed with
compressed air in the nozzle cavity. The compressed air jet is expected to enhance the kinetic energy of the
particles. The authors have reported about the Gaussian shape of the machined profile, with depth of around 100
nm within only 5 seconds of operation of N-BK7 Glass at 0.3 MPa. These preliminary tests proved that the
air/water combination of AJM can produce significantly more aggressive MRR with lower pressure. The increase
in abrasive’s kinetic energy is explained by the Venturi effect, although the particle to surface interaction within
the air-slurry mixture is not covered. Beaucamp et al. [41] undertook a task to control the number and size of the
air bubbles inside the jet. A novel system, which incorporates an ultrasonic transducer and specially designed
nozzle, was used to generate and manage the characteristics of air micro-bubbles through the frequency and
intensity of ultrasonic waves. Hence, removed volumes of both electroless nickel and BK7 glass substrates were
proportional to power growth but in reverse to a frequentative characteristic of the waves. The maximum
acceleration of MRR reached 380%, with even a slight improvement in surface roughness. A MRR dependence
on air bubble size was proposed, although the exact impact mechanism still requires further investigation.
2.7. Ice Assisted Jet Machining

Ice assisted jet machining (IAJM) uses grains of ice as the abrasive particles. The productivity of pure
water jet without abrasives is relatively low, however, the presence of crystallized water particles can drastically
increase MRR. Generally, IAWJ is an attempt to fill the gap between machining efficiency of pure water jet and
AWJM. Dry ice blocks can be integrated into water or the air jet through a standard method using a mixing tube
[43], however, the temperature of a system should also be controlled to ensure the survival of ice particles. [44].
Cryogenic wet-ice blasting with a temperature of around −100 °C was used to confirm temperature-dependent
hardness and removal capacity of ice [45]. The authors showed that the impact of cryogenic temperatures on the
ice particles must be evaluated critically since the ice with a temperature of − 60° C has shown higher MRR than
with −100 °C. Classical AJM, using mineral abrasives that are less environment-friendly, consumes unrenewable
resources and produces pollution. Since ice particles melt after processing, it can be used for cleaning of
contaminated surfaces. IAJM is also in demand for food [46] and medical industries as an effective cleaning and
disinfection method [47], where low processing temperatures are required to avoid bacterial growth. However,
the method has not been widely adopted due to the relatively high cost of dry ice and cold temperature maintenance
in IAJM system.

2.8. Intermittent Jet Conditions

Through eroding features with a high aspect ratio (deep grooves or holes), particles are often accumulated
at the bottom of the hole preventing the direct impact of oncoming abrasives with the substrate. Kuriyagawa et al.
[48,49] integrated a high-speed (100 Hz) solenoid valve into the blasting system, which periodically blocked an
abrasive flow to the gas jet. A continuous gas stream blows away particles that have accumulated on the bottom
of the hole. On one hand, a decrease of abrasive flow is expected to decelerate MRR, while on the other hand,
MRR is compensated by the substrate surface cleaning. Experimentation with 10/40 ms of open/close solenoid
time-cycle during 500 shots has proved an increase of hole depth from 440 μm to 740 μm. Time-cycle 10/10 ms
was almost inefficient, while 10/80 ms did not provide any growth of machining efficiency, in comparison to
10/40 ms. The method was applied to hole machining only, but its benefits for deep texturing in general are
obvious.

2.9. Submerged Jet Conditions

Most AJM methods are supposed to be operated in the atmosphere, nevertheless, some definite benefits
and new technological effects could be obtained when the jet is submerged below the fluid level in a tank.
Radavanska et al. [50] recommended submerged AWJ conditions for the sake of noise reduction, however, some
kinetic energy of the jet would be sacrificed. Shimizu [51] has shown that the submerged water jet in a stationary
slurry at definite pressure and SOD may cause cavitation erosion on the machined surface after 2 hours of
operational time. Due to increased drag on the submerged AWJ periphery, there was less diameter diffraction
along the jet [10,52]. Another argument for the application of the submerged process is a dust reduction. Abrasive
particles or chips penetrating into the lung could cause serious damage to human health, especially when
aluminium products are processed [53]. However, such working area pollution is fully eliminated when the
blasting process is submerged in a tank with liquid. Partial submerging of the process for 2 – 6 mm in liquid
glycerin, polymer, and water was adapted for AAJM with 61%, 42%, and 36% dust reduction, respectively [54].
The well-known weakness of all blasting methods, limited nozzle life, had been attempted to be resolved by
submerged machining conditions. Andilahay and Novikov [55,56] pushed the compressed air solely to the
workpiece submerged in the slurry. Submerged air flow is mixed with the abrasive particles beyond the nozzle
exit. This scheme eliminates the nozzle wear completely. The drawback of this method is the critically low particle
velocity (10 – 30 m/s).

2.10. Multi Energy Jet Conditions

AJM quality and productivity is closely related to the jet kinetic energy. An advanced solution for the safe
overcoming of the process energy threshold was adopted from the technological area of plasma machining. Cao
et al. [57] and Kim et al. [58] separated one powerful jet into a number of smaller jets, which allowed an increase
in the total power of the jets without significant degradation of its stability. Moreover, multi-jet showed
consistently better spatial uniformity than its single jet counterpart. The identical technique was applied to increase
fem-to-second laser irradiation [59] and enhance the production efficiency of EDM [60,61]. Wang et al. [62]
proposed naming this method “Multi Energy Jet Machining” (MEJM) and demonstrated its potential in enhancing
FJP efficiency without accuracy degradation. Resultantly, jet multiplication produced a proportional increase of
productivity with a surface roughness compatible with a single jet. The innovative rotational multi-jet nozzle was
reported by Shiou and Asmare [63]. A hexagonal tool head with six AWJM jets was directed at the machined
surface at 30° and designed eccentrically to avoid the interference between jets. Varying the number of jets from
2 to 6 with a difference in head rotational rate (20 – 60 rpm), Ra = 6 nm was achieved. MEJM is a promising
approach for machining of large area surfaces, where simultaneous operation of ten or more jets is possible.
However, due to nozzle wear, future increase of jet diameter and flow disturbance can potentially affect the inter-
jet interference and destabilize the footprint uniformity.

3. Capabilities and application

This chapter explains the AJM strengths and weaknesses, and demonstrates the diversity of
manufacturing operations in different industries where AJM approaches are applied. Achievements in
micromachining of regular patterns by AAJM and its application for tribological purposes are also analyzed.

3.1. Technological advantages and drawbacks

The diversity of AJM approaches provide various important advantages compared to conventional
machining technologies. Some of them are listed below:
• Manufacturing in the extensive range of machining scales, from macro [62] to micro [64]. The
process is almost boundless by the size and shape of the workpiece. The subtle water and magnetorheological jets
archive distant and hard approachable complex surfaces [28]. In addition, the jet may operate in the micro area
producing mild edge effect, that can be used for the edges and corners correction [65].
• Cutting process sustainability. Compressed air or water jet extracts the machining debris from
the cutting zone, creating flow conditions of chip removal on a majority of operations, except high aspect ratio
[48]. Abrasive jet is continuously re-sharpening its “cutting edge”. The process is not sensitive to the
nozzle/workpiece gap fluctuation due to vibrations or temperature changes [64].
• Process control flexibility. Machining spot, material removal rate, and surface roughness are
easily changeable through the numerous options of hydraulic, abrasive or machining process parameters.
• Slow tool wear with soft consequences. Gradual nozzle wear does not create abrupt distortions
to process accuracy for a prolonged operating time [5] in contrast to conventional cutting where cutting edge
radius is crucial. In most cases, AJM does not leave machining signature, moreover, cutting marks from prior
processes can be effectively removed [41].
• Improvement of metallic surface properties. Plastic deformation imposed during blasting yields
subtle sub-surface microstructural changes such as grain size refinement, martensite formation and work
hardening [66].
• Cost-effectiveness. Micro-abrasive blasting has minimal equipment costs and expenditure
during exploitation comparing to conventional milling, polishing and etching technologies [67]. Abrasives can be
recycled [68].
Despite the listed advantages, several process drawbacks are reducing its technological competitiveness:
o Tapered geometry of machined feature [19];
o A subsurface layer of brittle materials is impaired by micro-cracking [69];
o The productivity of micro-blasting does not exceed a few g/min [70];
o MRR during micro-blasting is unstable due to the poor consistency of fine particles flow [71];
o In AWJM, some operations are accompanied with striations and kerfs formation [72];
o In mask-less micro-patterning, machining spot grows due to gradual nozzle wear;
o Dangerous noise level (80 – 110 dB) [50,73] and risk of chemical intoxication [53,73];

3.2. Assignment and areas of application

AJM has almost ubiquitous manufacturing feasibility. Materials abrasion operation using non-fixed
particles can be implemented to the technological map of detail on many stages of the manufacturing process.
AJM can be applied to a coating [74] or welding as pre- or post-operation [75], providing removal of rust, scaling,
remnants of ingrained oil or outworn coatings. The productivity of micro-AJM approaches, except AWJM, does
not exceed a few g/min, for instance, up to 5 g/min in micro-blasting [70], thus it is not used for bulk formation.
On the other hand, a small machining spot is highly applicable for sheet material cutting, spatial surface polishing,
drilling, and patterning. At some conditions [76], surface strengthening can also be obtained as a result of the
peening effect [77]. In post-machining stage, AJM has gained increasing acceptability for deburring applications
[65,78,79] as well as for removing craters and cavities formed in the recast layer, for instance, after electrical
discharge machining [80]. List of operations provided by AJM is shown in Figure 4.

Technological operations of
Abrasive Jet Machining

Pre-machining
• Surface rust and scaling removal
• Surface oil and coatings removal
• Surface preparation for coatings
• Edges preparation for welding and soldering

Main machining
• Workpiece geometry fitting.
• Surface roughness reducing
• Surfaces material strengthening
• Micro-grooving and pitching
• Decorative surface patterning
• Matte or glossy surface obtaining

Post-machining
• Defective surface layer removal
• Local surface defects removal
• Deburring and rounding off workpiece edges
• Seams machining after welding

Figure 4. The functionality of Abrasive Jet Machining.


Figure 5. Polishing-turning operation with AWJM [81].

Due to its unique technological capabilities and broad functionality, AJM is used in machine building,
metallurgy, and other industrial sectors. In manufacturing, AJM performs the functions of cutting, polishing,
milling, drilling and surface texturing, superseding such conventional tools like grinding wheel, lathe cutter, mill,
drill, etc. as shown in Figure 5. In metallurgy, the abrasive jet is used for the cleaning of cast blanks from moulding
and baking. In electronics and instrumentation, micro-blasting is utilized for manufacturing microfluidic devices
[82] and a new generation of precise optics [30]. In civil engineering, sandblasting is considered to be one of the
most effective techniques for rust and graffiti removal [83,84]. Ice jet machining supplants a “kitchen knife” in
food [46] and medical industries as an effective cleaning and disinfection method [47], where low processing
temperatures are required to avoid bacterial growth. AWJM was adapted in the textile industry for cutting and
cleaning of garments materials [85]. Decorative patterning and drawing of images on coloured stones and glasses
using stencils promotes utilization of AJM in art. Due to the absence of thermal damage and consistent sharp
“cutting edge”, micro-AJM has the potential application as a substitute to solid instruments in orthopaedic surgery
[86–88], plastic surgery, neurosurgery [89], dermatology [90], urology and medical implants manufacturing [91].

3.3. Texturing capabilities

Surface patterning is a process of modifying surface topography in order to create a uniform microrelief
composed of regularly distributed asperities or depressions with controlled geometry [92]. Application of AAJM
for texturing purposes have shown a constant positive trend, starting from 2000 [82,93,94] when blasting with
fine abrasives was proved to be a high-quality mechanical etching technique. Sandblasting, which was initially
designed for rough working such as deburring and rough finishing [73], took a course to process miniaturization
in a new century. Higher operational resolution and finer surface patterns were promoted by the manufacturing
needs of semiconductors, electronic devices and LCD [95]. In literature, AJM and blasting are accompanied with
the annexe “micro”, e. g. [74,96–99]. The fair border for AAJM and micro-AAJM would be the orifice diameter
of 1 mm. Since 2000, more than a half of AAJM investigations have dealt with nozzles below 1 mm of inner
diameter.
Patterning resolution at the micro-scale was mastered in 2005. Slikkerveer et al. [93] machined 500 –
100 μm wide channels in glass by 23 μm particles with erosion depth homogeneity of within 5%. They concluded
that the dimension of a pattern is limited by the mask wear and edge damage, predicting minimum feature size of
around 10 to 20 μm. A few years later, 80-μm wide channels in glass were presented by Park et al. [95] and at the
same time, 50x50µm pockets for micro mould dies in stainless steel [100]. An extreme feature size – 10-μm wide
channels, which are still unique, were achieved by Wensink [101] (see Figure 6).
The first obstacle for patterning in the lower end of micro-scale is a suitable mask material which is
capable of defining such feature size. Wensink [101] had proved an excellent erosion resistance and high
suitability of electroplated copper for 10-μm mask openings. Meanwhile, photoresist SU8 and Rayzist masks
allow the minimum feature size of 75 μm and 80 μm, respectively [102]. Under this limit, the mask structures do
not open anymore, and abrasives were impeded to hit the target. The second obstacle is an application of fine
particles, as small as 3 µm, that are connected with the flow consistency problem. Neglecting vibratory feeding
systems, a powder flow repeatability problem can be observed even with particles as large as 25 µm [71]. Also,
with such a small particle, due to the low kinetic energy, brittle fracturing transits to ductile removal mechanism
[69]. This leads to a very bad selectivity between the mask and the substrate, which makes it difficult to machine
100 µm channels with aspect ratio deeper than 2.5 [101]. The third is the minimum crack length in the target
material. According to the erosion model [69] the smallest lateral crack, for example in Sola-lime or Pyrex glass,
is around 6 μm, that can significantly impair the edge geometry of the 10-μm wide channel. From this position,
particle energy should be lower than the energy threshold for lateral crack origination, which turns erosion in
ductile mode and improves the edge geometry of the channel but causes the reoccurrence of the mask selectivity
problem.
Following that, the chase for minimum feature size had been subsided. Further investigations were
focused on footprint profile predictability, in which the feature size was barely descended below 200 μm. Analysis
of blasting profile development was presented in the works of Achtsnick et al. [103–106] and furthermore in
analytical-empirical [107–112] and numerical [113,114] approaches of Papini`s group. Kef and holes typically
have tapered geometry. Oscillation of the substrate [115], continuous change of SOD [116] and different
combinations of process parameters [117] were applied to ensure steep sidewall slope and flat bottom of the
machined footprint.

a b C
Figure 6. Micro-blasting texturing resolution: a) 500-μm wide channel profile blasted by two cross nozzles in
glass using 30-µm particles [93]; b) 70-µm wide channels in silicon using 9-μm particles [101]; c) 10-μm
wide channel blasted in glass using 3-μm particles [101].

3.4. Outlook for tribology

Surface texturing capabilities of micro-AJM are often stereotypically associated with microsystem
fabrication. In this paragraph, we demonstrate a poorly recognized but powerful prospect of AJM micro-patterning
facilities for tribological purposes, using an example with extremally restricted requirements – artificial implants.
In the 2000s, around half a million hip and knee replacements were performed in the United States alone
[118]. Four million are anticipated annually in 2030 [119]. Up to 80% of all the surgical revisions of joint
prostheses must be performed due to loosening of the implant caused by sliding wear [120]. Wear resistance is
associated with material hardness, surface topography and integrity [121,122]. The most successful replacements,
yielding 90% success rate over 15 years period, are based on polyethylene-on-metal friction pair [123]. Owing to
the low thermal resistance of polymers, such machining methods like grinding or lasering produce thermal damage
and thus are inappropriate. Conventional machining of metals is characterized by the presence of surface defects
like cracks, scratches, cavities, burrs, etc. Indicated surface imperfections are responsible for polyethylene wear
acceleration, and consequently for bearing life shortening of about one order of magnitude [124,125]. Another
problem occurs during exploitation: bone cement debris left in the articulating surfaces during surgery or metallic
carbides debris formed in the contact area during articulation. Both produce damage and accelerate the wear rate
[126]. A specific topography of frictional surface (see Figure 7 for example) can serve as wear debris collector
and exclude its contribution to surface abrasive wear [127]. In addition, collectors may store and supply lubricant
to the frictional surface, mitigating the wear rate [128]. It was proved that micro-patterns are proficient in
improving the working period of an artificial joint over more than 30 years [127]. However, laser texturing is
pernicious for prosthesis due to the presence of thermal damage, such as melting and flow and severe bulges and
burrs [129]. Laborious CNC micro-drilling has relatively low texturing performance and is fraught with burr
formation on the hole edge [130]. Electrochemical and photochemical texturing methods are inapplicable to
polymers [92]. The application of the most prospective material for artificial prosthesis – ceramics [123] is
stagnated due to the presence of the damage layer after conventional surface finishing [131].
Applying AJM to both soft and super hard materials, surface roughness in nanoscale and the undamaged
superficial layer can be provided (see chapter “Material removal mechanism”). Gradeen et al. [36,37] switched
ductile removal mode to brittle and improved the surface roughness of polymers using cryogenic temperature
conditions during AJM. Wakuda et al. [8] reported an absence of strength degradation in ceramic surfaces after
AJM, concluding that the radial cracks propagation downwards to the material might be eliminated. Therefore,
AJM has a high potential as a damage-free machining method for a wide range of materials. Nakanishi et al. [132]
have used micro AWJM to polish (Ra = 1 nm) and engrave (1-mm wide groves) the surface of Co-Cr-Mo alloy
by means that reduced the wear rate of prothesis fourfold in comparison to a conventional mirror finish surface.
Further investigations have shifted the feature sizes down to 100-μm diameter dimples [133] and 40-μm wide
grooves [134]. The trend illustrates that for the life prolongation of artificial joint and osteolysis prevention, the
size of micro-pattern is moving to the lower end of the micro-scale. As was presented in the previous section, the
masked micro-AAJM texturing technique allows fabricating channels as narrow as 75 μm [102], 35 μm and even
10 μm [101]. Such capabilities create a powerful prospect for AJM in tribology, realizing required regular, half-
regular or non-regular patterns on the freeform surface of any hardness.

Figure 7. Cobalt-chrome alloy head (D = 26 mm) with concave pattern (d = 0.5 mm, h = 0.1 mm)
[135].

4. Material removal mechanism

Erosion is conventionally considered as a negative phenomenon, producing damage to structures. In the


conception of free abrasive machining, erosion becomes an instrument, where AJM is a manufacturing
technology, which is based on erosion localization and intensification. Directed flow of hard micro-particles
splits-off the tiny chips from the substrate, removing workpiece mass to required geometrical conditions.
Depending on material’s properties and process parameters, ductile or brittle removal mechanisms may dominate
during erosion. The third one, namely “elastic removal mode” was discovered recently. The understanding of
material removal mechanism is essential for machining efficiency enhancement.

4.1. Brittle Material Removal Mode

Numerous studies were devoted to an explanation of MRM in brittle materials [69,136–139] after directive works
of Marshall, Lawn, and Evans [140,141]. Briefly, the deformation and cracking model that occur during particles
impact with a surface are typically those known from quasi-static Vickers-indentation theory [141]. The ideal
cracks pattern is represented schematically in Figure 8Error! Reference source not found.. Particles indentation
creates compressive stress in the material beneath that forms a plastically deformed area. When the fracture
threshold is exceeded, a radial crack perpendicular to the surface propagates downwards and aside from the base
of the deformed area. During unloading stage, the lateral crack occurs at bottom of the plastically deformed
volume and extends parallel to the surface. The radial crack does not affect the chip formation, nevertheless it
degrades the surface integrity. It is generally accepted that the lateral crack determines the removed volume,
assuming the chip size as a hemisphere with a volume dependent on lateral crack radius and depth of origination.
Correspondingly, the radius of lateral crack and the depth of its initiation are considered to be the explanation of
erosion phenomenon.
Buijs [136] calculated the radius of the deformation zone and length of lateral crack from the position of
indentation-fracture mechanics. These values are the functions of viscoelasticity, hardness and fracture toughness
of the workpiece, particle geometry and indentation force. During blasting of ceramics, the cracking mechanism
was well correlated with the classical indentation model. However, due to different levels of ceramics
inhomogeneity, it was impossible to generalize the model for each individual material. The same author, referring
to previous studies, proposed the expressions for equilibrium length of the lateral crack, the bottom of the
plastically deformed zone and the volume removed by single particle indentation. Later, a threshold energy for
radial and lateral crack origination was found by Slikkerveer et al [69]. According to the authors, there is a
minimum length of crack for each material. If the kinetic energy of the particle is lower than the energy required
for minimum crack formation, then the material response to particle`s impact is elastic/plastic. Jafar [137] assumed
that lateral cracks initiated at the bottom of the indentation depth rather than at the bottom of the deformed zone.
This assumption allowed improvements in erosion and surface roughness predictability threefold. However, the
physical reason of lateral crack initiation from the indenter tip remains unclear. The weakness of radial/lateral
crack systems lay in the assumption of isotropy of material structure. Wakuda et al. [8] blasted sintered ceramics
and showed that the multiple ceramic grains were fragmented simultaneously in a form (Figure 9), which
substantially differs from the assumed hemisphere crater in the indentation erosion model (Figure 8). The
conclusion that the microstructure is the primary factor in the erosion behaviour of ceramics is supported by [142].
Thus, the lateral/radial crack propagation theory, based on the assumption of material`s homogeneity, cannot
entirely expound the erosion mechanism in brittle materials.

F (Ukin) Substrate
Abrasive
particle Chip
Compressed
zone Depth
of crack

Lateral crack Length of crack


Radial crack

Figure 8. Schematic crack propagation in a result of particle indentation.

Figure 9. Fragmentation from the single impact of SiC particle with Al2O3 substrate [143].

4.2. Plastic Material Removal Mode

The primary ductile removal mechanism at normal impact is still under discussion. In contrast to shallow
impact angles, where material removal is well explained through the cutting process, ductile erosion at normal
impact is a polyhedral combination of process kinematics with dynamic material properties. After a final summary
of the basic mechanisms by which solid particles may remove soft material pointed out by Finnie [144], no great
fundamental clarifications were introduced. The low-cycle fatigue mechanism was used to introduce “erosion
ductility” [145], which is conceived to be an as essential material characteristic responsible for the critical plastic
strain. Another erosion parameter was formulated considering the energy criterion during erosion and incorporates
the material’s hardness, along with the “high-strain-rate stress-strain response” of the substrate [146]. However,
these approaches are targeted on quantitative evaluation of volume removal and shined little light on the physical
mechanism of material removal. A new elastic-plastic analysis was proposed [147] to predict the crater parameters
and rebound velocity that can be used to calculate the amount of plastically deformed material appearing at the
edges of the impact crater. It is known that the deeper into the crater, the more resistant the substrate becomes.
The work of hardening during impact is very difficult to evaluate. Dynamic hardness widely applied in erosion
models, e.g. [145,148,149] reveals to be more of a functional value rather than a material property, depending
greatly on erosion conditions such as particle material, size, and velocity, coating thickness, etc. For instance,
dynamic hardness of PMMA may vary from 970 MPa [150] to 2600 MPa [151] depending on testing conditions.
It is unlikely that current micro-blasting conditions (3 – 30-μm particles) can be reliably imitated in dynamic
hardness testing.
According to concepts of ideally ductile removal mode, such mode does not leave cracks and thus is
preferred over brittle mode for machining of solids (e.g. optics, microfluidic devices) where surface integrity is
essential. Despite the common expectation, ductile cutting was observed on brittle materials at large impact angles
[142]. It was explained [143] by the fact that local instant temperature at the impingement spot may rise up to
1500 K [152], which reduces material hardness and allows it to realize the ductile-brittle transition, for example
on alumina ceramics [153]. Another approach to shift erosion from brittle to ductile mode is to reduce the kinetic
energy of the impact particle below the threshold energy of crack generation [69]. For brittle materials, the ductile-
brittle transition takes place from 6·10-7 J to 6·10-10 J [154]. However, again, after critical strain accumulation,
the crack may occur. It should be remembered that the concepts of ideally ductile and brittle behaviour are just
useful approximations.

4.3. Elastic Removal Mode

Material removal in ductile or brittle mode, based on deformation and consequent crack propagation,
inevitably lead to a change of sub-surface structure. Surface integrity mostly depends on particle size [39]. Finer
grains produce smaller plastic flow and depth of impaired surface. Reduction of particles size from 120 nm down
to 40 nm still have an influence over MRR, although produce no statistical effect on surface roughness [155]. In
other words, material removal with nanoparticles may not be related to deformation and cracking. It was
concluded that the removal mechanism was shifted from indentation mechanism to “surface-area mechanism”.
The particles trajectory model [156] demonstrates that the abrasive particle with a diameter less than 100 nm,
turns away with the fluid flow just above the collision surface. Relatively large grain of more than 5 μm move in
a straight line neglecting the turn of fluid flow and impacts the workpiece at almost initial impact angle. Such
particle behavior is explained through the significant domination of the centrifugal force over the Stokes resistance
while enlarging the physical size of the grain. At a definite ratio of particle mass to velocity, the normal force
gradually turns to tangential and particles touch the substrate almost asymptotically (see Figure 10, particles with
diameter d2). In the case where impact force is less than the threshold required for material deformation, merely
an elastic interaction takes place. It was proposed by Peng et al. [156–159] to provide critically small contact
forces and remove the material through the chemical impact reaction between particle and substrate. Chemical,
or as proposed “elastic” material removal mode, is based on the surface hydroxylation effect and chemisorption
theory [160–162]. When nanoparticles overcome the energy threshold under certain impact activation, the atoms
allocated on the surface of nanoparticle would chemically connect to the substrate surface atoms to form a covalent
bond with the surface distance about 1 nm. For instance, during glass polishing, the following chemical impact
reaction would occur [163]:
[SiO4 ]n Si(OH)m + (OH)m R → [SiO4 ]n SiOm R + mH2 O

where, R – identify the atoms of abrasive particle, m – hydroxyl number of glass surface silicon atom. Silicon
atoms are connected to particles hydroxide and removed from the glass surface by the flow shearing action, as
demonstrated in Figure 11. In one of the experiments, Peng et al. [156] smoothed the surface roughness down to
0.178 nm in Ra. Plastic scratches and pits obtained from the previously processed surface were removed away.
Since the arrangement of each atom recovered back to its initial position, elastic interaction did not produce any
defects to final surface integrity.
Particles
with diameter
Impact
direction

Substrate surface
Figure 10. The trajectory of different size particles at the workpiece surface [156].

Figure 11. Chemical impact reaction as removal mechanism [157]

5. Influence of process parameters

AJM process is affected by the number of settings. Some factors may contribute differently depending
on the combination of other factors and materials properties. Although, several dominating tendencies can be
underlined. The independent process parameters involved in AAJM were classified by Hashish [164] into two
general groups and later into three groups by the Nouraei et al. [13], which are discussed below.

5.1. Hydraulic parameters

5.1.1. Nozzle pressure

Pressure directly affects flow velocity and, as an aftermath, the kinetic energy of the in-flow particles.
Thus, an increase in pressure leads to the growth of MRR and surface roughness. In AAJM, working pressure
may vary from 0.2 to 1 MPa, that usually corresponds to 100…300 m/s of particle velocity. During pre-coating
machining for improvement of coated layer bonding to zirconia substrate, a working pressure less than 0.2 MPa
was recommended [165]. For post-coating treatment of TiAlN, the optimal nozzle pressure in terms of roughness
and MRR was given between 0.2 MPa and 0.4 MPa [97]. The highest hydrophobic properties of
polytetrafluoroethylene were observed after blasting with 0.6 MPa [166]. For surface smoothening, relatively low-
pressure conditions are generally concluded (Figure 12). Low pressure is also reported to produce steeper kerf
slope [117].
1.6

Surface roughness, μm
1.4
1.2
1
0.8
0.6
0.4
0.2
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Nozzle pressure, Mpa
PTFE substrate, NaHCO3 particles of 270 μm,
stand-off-distance 53 mm
TiAlN substrate, Al2O3 particles of 100 μm,
stand-off-distance 100 mm
Y-TZP substrate, SiO2 particles of 110 μm,
stand-off-distance 10 mm

Figure 12. Nozzle pressure influence on surface roughness in AAJM (data collected from [97,165,166]).

5.1.2. Jet velocity

Due to boundary conditions between air jet and stable surroundings, abrasive jet velocity is non-uniform
in both radial or axial direction. Li et al. [167] conditionally divided air flow into three velocity regions. In the
initial region, jet velocity preserves its speed in a form of conus with a base at the nozzle exit. The highness of the
conus (“Core” on Figure 13) is independent of the jet velocity at the nozzle exit. Instead, it is proportional to its
diameter and flow viscosity. Rajaratnam et al. [168] informed that the length of initial region is around 6.2 times
of nozzle diameter for air jet and near to 100 times for water jet. The last one – impact or diffused droplet region
is the area of abrupt flow deacceleration and turning into parallel to workpiece surface direction. Inside potential
core, air velocity is higher than that of the particle, thus the particle preserves its acceleration until it becomes
faster than air. Maximum particle velocity is achieved near to the beginning of main region [98,169]. Further, the
centre line particle velocity decreases almost linearly as the distance from the nozzle exit increases [170].

dn

Ln Nozzle

Initial 195
Lc Nozzle
Core region Core
Air flow velocity,

diameter
760 μm

98
Main
m/s

Lm region

Div.
0

Impact
region

Figure 13. Scheme of the conditional division of jet velocity regions (left)[13] and microscope images of the
AJM jet (right)[171].
5.2. Machining parameters

5.2.1. Traverse speed

Traverse speed is a speed of the reciprocal motion of the blasting nozzle relative to the machined surface.
The speed selection is based on requirements of feature geometry. In precise etching by micro-blasting, the speed
may fall down to 0.25 mm/s [71,107]. The slower speed provides deeper erosion spot. Particle distribution in a
flow cross section has normal character [104], consequently, the abrasive jet produces Gaussian shape footprint
with a bottom at the distribution centre [114]. However, at some conditions, the machined profile can turn to the
flat or even convex pattern, as was reported recently [117].

5.2.2. Feed step

Feed step is a length of nozzle axis shift for each path alongside the previous one. Correct sequential
groves overlapping with a small erosion depth may be considered as polishing. With an eye to provide a small
pick to valley value of machined surface, the feed step should be small enough to conjugate the flat regions of
two bottoms, otherwise, sinusoidal surface profile occurs (Figure 14).

Figure 14. W-shaped channel in glass [172]

5.2.3. Stand-off-distance

Stand-off-distance is the distance between the nozzle exit and machined surface. Moving from
conventional sand-blasting to powder-blasting, SOD was shortened from a few meters to several hundreds of
micrometres. There are two effects of SOD: a) increase of SOD above initial jet region leads to a reduction of
particle velocity [167], reducing both MRR and surface roughness; b) increase of SOD enlarges footprint width
due to natural jet divergence. In AAJM, particle divergence angle is typically around 10° as noted by Ghobeity et
al. [108]. However, at low pressure, around 0.2 MPa, divergence can be minimized down to 1.5° [171]. Owning
to CFD model of Kowsari et al. [171], the enlargement of the footprint is also connected to the effect of particles
secondary impact. The kinetic energy of rebounded particles at the second impact can be large enough to erode
the substrate again. The larger the SOD, the broader the trajectory for the second impact (Figure 15). The second
strike area grows or falls when the convex or concave surface is machined, respectively [173].
Figure 15. Particles trajectory at first- and second-strike [171].

5.2.4. Incidence angle

Incidence or impact angle is an angle between nozzle axis and machined surface. Depending on the target
hardness, nozzle inclination causes different aftermaths. Large angles (close to 90°) provide higher MRR in brittle
materials [108], particularly in ceramics [142], while the soft workpiece is cut more efficiently with the angles at
around 20° - 30° [107], as shown in Figure 16. The normal force component is dissipated to plastic deformation
and to the propagation of intergranular microcracks, resulting in grains detachment. The tangential component of
the impact force is mainly for cutting and ploughing. The presence of hard metallic grains in ceramic structure
prevents it from cutting effectively. In an experiment [142], the erosion rate of alumina ceramics at 15° impact
angle was 5 times lower than at normal impact. On the other hand, owing to a minor normal component of impact
force, there was smaller imprint depth and hence, smoother surface. Since the abrasive flux is spread over a wider
surface area, it also results in footprint geometry, giving an asymmetrical form [109].

1.2 Grain Material Vickers Ref.


Normalized erosion rate

size, Density, Hardness,


1 μm g/cm3 kg/mm2
0.8 PMMA 1.2 250 [107]
316L SS 8.03 1373 [52]
0.6
Al 6061 T6 2.7 981 [52]
0.4 Al2O3 <1 3.88 1800 [174]
0.2 AD998-C 12.7 3.88 1720 [142]
AD92 7.5 3.61 1340 [142]
0
0 10 20 30 40 50 60 70 80 90
Impact angle [ged]

Figure 16. Normalized erosion rate as a function of impact angle for polymethylmethacrylate, stainless still,
aluminium alloy and three alumina ceramics.

5.3. Abrasive parameters

5.3.1. Abrasive size

Abrasive size is one of the most influential factors in any AJM modification. By increasing the particle`s
size, the single grain obtains bigger mass and volume, that directly affect its kinetic energy, which in turn
influences MRR. An increase in particle size is limited by the nozzle diameter. Particle’s interactions within the
stream can reveal a problem, even when the particle diameter is 15 times less than that of nozzle [148,175].
In general, the increase of abrasive fraction leads to chips enlarging and surface roughening [165].
However, machining the sintered ceramics using 25-μm and 10-μm alumina particles at the same conditions,
Kowsari et al. [174] reported an approximately identical peak-to-valley roughness. In contrast to homogeneous
materials such as glass, in sintered materials, the erosion occurs through the detachment of individual grains and
thus tightly connected with the grain size of the metallic structure. Surface smoothening through the refining of
abrasive fraction is also limited by the physical size of particles. Smaller particles possessing a higher square to
volume ratio are very sensitive to humidity level in a storage. Adsorption of moisture enhances interparticle
adhesion [176,177], affecting abrasive flow consistency [71].
Another factor potentially affected by abrasive size is an embedding phenomenon, particularly common
in soft substrates during AAJM [36,150,178]. It is obvious that embedding is proportional to particle kinetic
energy. Although, even in FJP [41] at a slow impact velocity of 10 – 70 m/s, alumina particles between 0.6 and
18 μm were found plentifully embedded to Tungsten carbide coating (Figure 17). Getu et al. [150] concluded that
the likelihood of embedding is proportional to the static coefficient of friction between the particle and the
substrate. Static coefficient of friction for angular particles was found to be independent of particle size and impact
angle for the investigated particle/target systems. In experiment [37], all of the particles found embedded were
significantly smaller than the nominal particle diameter. Hadavi et al. [179] measured and modelled an
instantaneous particle orientation and concluded that larger particles are more likely to orient themselves in the
jet direction than smaller ones and thus, are more likely to embed. Clearly, the influence of particle size on
embedment should be discovered.
Concerning the machining spot, fine abrasives are perceivable to air flow turbulence and thus impact the
substrate at a broader area. More massive abrasives, possessing a higher centrifugal force, overcome the lateral
air flow and interact with the surface at an almost initial angle [156]. The increase in masked channel width with
decreasing particle size is a known phenomenon [110], which is due to the fact that smaller particles can pass
closer to the mask edge without striking the mask. By shifting particle size from 10 μm to 25 μm, 9% of channel
narrowing was detected [180].

Figure 17. Particles embedding dependence on its size and fluid jet pressure [41].

5.3.2. Abrasive hardness

Abrasive hardness directly exerts to MRR. Wakuda et al. [8] blasted various ceramic substrates by
abrasives with comparable angular shape and size, but different Knoop hardness. As it can be seen from Table 1,
in the case of ZrO2 substrate, the erosion rate was almost independent of the abrasives hardness, such as workpiece
hardness which was substantially lower than the hardness of any abrasives. For harder ceramics, the erosion rates
were noticeably less for aluminium oxide and silicon carbide particles, although, not for synthetic diamonds. It
was concluded that when the hardness of abrasives is close to substrate`s hardness, some part of the kinetic energy
may be converted to deformation, heating, and fracture of the particles itself. On the other hand, when superhard
particles were directed to the soft substrate, above described embedding phenomenon may take place.
Surface roughness also has direct dependence on abrasive hardness. Alumina particles are not hard
enough against the ceramic substrates to seriously engrave its surface. Wakuda et al. [143] explained this
smoothening through the transition from brittle to ductile removal mode due to a high local temperature at the
impact area [26-27]. In the case of superior hardness of synthetic diamond, there is no significant heat energy
released during impact, such as a rough surface with large-scale brittle fragmentations was observed. Hence,
application of superhard abrasives might serve for high machining efficiency, although conditional for surface
smoothening.

Table 1. AAJM footprint for various abrasives and ceramic materials settings [8].
Material properties of the tested ceramic substrates
Ceramic alloy ZrO2 Si3N4 Al2O3 SiC
Density (g/cm3) 6.05 3.2 3.9 3.1
Young’s modulus (MPa) 210 290 390 390
Vickers hardness HV (GPa) 13.2 14.2 15.3 22.1
Fracture toughness KC (MPa/m) 7.0 7.5 4.2 2.5
Flexural strength (MPa) 1200 1000 360 470
Aluminium oxide
Mesh size 800
Density (g/cm3) = 3.9
Knoop hardness (HK) = 2100
Abrasive particles

Silicon carbide
Mesh size 800
Density (g/cm3) = 3.1
Knoop hardness (HK) = 2480

Synthetic diamond
Mesh size 800
Density (g/cm3) = 3.5
Knoop hardness (HK) = 7000

5.3.3. Flow rate

Abrasive flow rate is the mass of abrasive powder supplied to a mixing chamber per unit time. Typically,
flow rate varies from around 1 g/min [110] to 1 kg/min [166], depending on nozzle diameter and pressure. For
micro-AAJM, it is typically up to 5 g/min. In water-based AJM, flow rate may be expressed in particle mass per
volume of liquid, for instance, 40 g/l [40], or as a percentage of particles concentration in a slurry, where a range
from 0.25% [13] to 10% [181] is usually investigated. Cylindrical nozzles have a Gaussian profile of particle
distribution across the jet as well as Laval nozzle, however, the latter has a more homogeneous distribution [103–
105]. However, Li et al. [117] demonstrated that flow dynamics is more complex and related to process
parameters. Reduction of abrasive flow rate allowed the transformation of a well-known U-shaped machined
profile to a W-shaped profile. In their opinion, the air stagnation and particle bounce back effects can cause the
distribution to shift from center to brim concentrated pattern, depending on the particle flow rate and pressure.
The hypothesis is supported by particle deposition experimentation [182], although the conditions there were far
different from those in micro-blasting. A further study is needed to understand and evaluate the effect of particle
flow rate on its cross-sectional distribution. According to the mathematical model of Fan et al. [183], the MRR is
inverse to flow rate, such that additional particles deaccelerate fluid flow and reduce the amount of kinetic energy
distributed to a single particle. However, at the high-pressure level, when all additional particles have sufficient
velocity for substrate fracturing, or when a growth of abrasive flow rate is accompanied by the parallel rise of
pressure, process productivity (volume removed/time) increases. Particles may pack together and form local
cavities in the powder reservoir which may produce significant variation in the particle mass flow. Such variations
alternate the MRR and channel depth, respectively. Vibrational feeders with in-suit powder dryer (e.g. Texas
Airsonic, HP-system) demonstrated performance with 3 – 9 μm particles [69,93,101], however, flow reputability
was not measured. Unfortunately, no reports on performance with such small particles are available in literature
for air-pressurized powder feeding systems. In order to ensure the best flow consistency, the following
recommendations were given [71]:
➢ The use of a powder reservoir-mixing device;
➢ The low powder level in the reservoir;
➢ The use of desiccant and refrigeration-based air dryer;
➢ The placement of packets of desiccant in all powder storage containers;

5.4. Process parameters summary

The synopsis of given chapter is presented in Figure 18. The table reflects general tendencies and may
not correspond to a specific case. To avoid misconceptions, a few comments should be added. Particle velocity
grows for some SOD and then starts to fall [167], which consequently affects the amount of kinetic energy, MRR,
roughness, etc [69]. Impact angle has a different affect to MRR, depending on substrate properties [184]. The
increase of flow rate reduces particle velocity, its kinetic energy and volume removed in each impact, but even
so, it improves process productivity (mass removed/time). Since micro-texturing is usually performed with the
use of masks, the influence of particle characteristics on footprint width was out of the researchers focus. Particle
velocity and kinetic energy are independent of its hardness or angle of impact.

Particles
velocity

Particles
kinetic
energy
Material
removal
Function

rate
Surface
roughness

Footprint
width

Footprint
depth

Independent parameter
Nozzle Stand-off Incidence Abrasive Abrasive Abrasive
pressure distance angle size hardness flow rate

Figure 18. Influence of process parameters. Synopsis.

6. Nozzle wear issue

6.1. Wear mechanism

Like most of the other machining technologies, all AJM methods are related to the issue of tool wear.
The nozzle is the most vulnerable component of any abrasive jet system. The typical working scheme of the nozzle
with a mixing chamber is presented in Figure 19 [138,185]. High pressured energy carrier moves through the
orifice to the inner chamber, where it is mixed with abrasive particles. Then, the mixture enters the nozzle tube,
obtaining a directed motion and exits in a form of an abrasive jet. The wall of the mixing tube is multiply impacted
by particles during the process, that leads to internal nozzle erosion and changes in the tube profile. Continuous
increase of nozzle hole diameter leads to process instability due to rise in air flow rate, jet divergence, and footprint
size. Such circumstances affect the MRR, surface waviness and preciseness in general. Therefore, the nozzle wear
mechanism became an important technological topic in the improvement of AJM economic indicators.
Compressed air

Tube from
compressor

Orifice

Abrasive
flow

Mixing
chamber
Case

Nozzle

Abrasive Air Jet

Figure 19. Typical constriction of AAJM nozzle with mixing chamber.

During blasting, abrasive particles collide with the entrance edge and inner cylindrical surface of the
nozzle as shown in Figure 20. It leads to erosion and geometry changes both on the nozzle entrance and exit. A
common nozzle life indicator is nozzle weight loss rate (NWLR) in g/min. Depending on impact angles, both
ductile cutting and brittle fracturing may proceed simultaneously [102, 152, 155–157]. It was shown that the
central and exit sector of the nozzle undergoes mainly the ploughing with a linear wear rate, meanwhile, there is
a brittle fracturing at the nozzle entrance [186]. Boron carbide nozzle has shown the rapid wear for the first 10
hours of operational time and then slow progressive wear for the next 50 hours. Alumina nozzle with the same
geometry undergoes rapid wear for the first 4 hours [138]. Both cases demonstrate the same NWLR character
which makes relevant the conclusion that the sliding erosion mechanism after some operational time starts to
dominate over the fracture produced by the direct particles impact. The nature and rate of nozzle wear is affected
by a number of factors, which could be separated into three groups: nozzle material, geometry, properties of
abrasive particles and influence of process parameters, which are discussed below.

Impact at 90o New nozzle

Fracturing Shallow impact

Ploughing Worn nozzle

Figure 20. Schema of nozzle-particles interaction [138] and worn nozzle photo [187].

6.2. Influence of nozzle material

A variety of ceramics, especially silicon carbide (SiC), alumina oxide (Al2O3), zirconia dioxide (ZrO2),
boron carbide (B4C) and silicon nitride (Si2N4) have got an extensive application in nozzle manufacturing due to
its universal mechanical properties. In comparison to metals, ceramics have a higher hardness and melting point,
but lower thermal shock resistance. Lastly, significant thermal stress during the blasting process results in
intensive brittle fracturing at the nozzle exit [188]. Nevertheless, for instance, alumina ceramic nozzles are able
to work for around 30 times longer compared to stainless and carbon steel nozzles [189]. Thus, hardness was
claimed as one of the most important contributors to the nozzle life. The hardness and abrasive resistance of boron
carbide are inferior only to diamond and its relative cubic boron nitride structure. Monolithic B4C coating shows
a reduction of NWLR for around 30% in contrast with Al2O3/(W,Ti)C coating [138]. Sapphire orifice was used
for ultrasonic CAFJP [40,190] and for the creation of "hydraulic flip" effect [191].
In studies [139,187,192–195], functionally graded materials (FGMs) in a range of two ceramics were
applied to nozzle design. FGMs is an approach to design parts with gradually changing properties by means of a
sequential combination of different materials. Comparing to sintered (W,Ti)C, SiC has a low coefficient of thermal
expansion and high thermal conductivity. To reduce tensile stress and wear, (W,Ti)C nozzle was gradually
modified to SiC from the nozzle exit to nozzle entrance (Figure 21). FG nozzle has shown almost 4 times less
NWLR than the conventional one. The physical reason for such reduction was demonstrated by means of finite
elements modelling of nozzle stress. FE model of FG nozzle reveals significantly less stress at the nozzle entrance
compared to conventional design, 129.11 MPa, and 521.58 MPa, respectively. FG nozzle with gradation from
Al2O3/(W,Ti)C to Al2O3/TiC shows an improvement in fracture toughness. The difference in thermal conductivity
coefficients ensures a reduction of temperature gradient and with this, decreases the thermal stress.

Figure 21. Functionally graded nozzles [195]

6.3. Influence of nozzle geometry

Increase of nozzle length decelerates the wear rate at the nozzle exit. It was explained [185] by the
vector`s components of the particle velocity. Elongating the nozzle, the particle velocity vector becomes more
parallel to nozzle axis, losing the perpendicular component responsible for particles indentation to nozzle wall.
The tool life of a 76mm long nozzle was doubled by elongating to 127-mm. The dependence is not linear, thus, it
is the subject of optimization.
Nozzle diameter growth results in NWLR falling. The ratio of the orifice to nozzle diameters for the
optimum mixing and cutting conditions is close to 0.4 [196]. Nozzle weight loss descents when approaching this
value from below and slightly accelerates when it overcomes.
Unequal wear along the axis of the nozzle is contributed by different tensile stress arrangement and local
thermal expansion. Maximum tensile stress occurs at the entry edge of the nozzle tube due to direct particles
impact. Deng et al. [186] proposed to enlarge the hole entrance under definite slope. The nozzle entrance angle of
11◦ was found to be the optimal, further expansion of angle led to gradual stress growth. As the inlet angle
increases, nozzle profile becomes wavier, resulting in increasingly non-linear growth at the exit. Smaller inlet
angles show better linearity in the character of nozzle exit enlargement [197].

6.4. Influence of abrasives and process parameters

Increase of abrasives size and hardness lead to the almost linear growth of NWLR. During AAJM with
SiC particles the cumulative mass loss of B4C nozzle was doubled when abrasive fraction of 180 – 212 μm was
replaced with 250 - 300 μm [198]. Finer particles have minor kinetic energy and consequently produce smaller
impact loading. Application of 33 GPa hardness SiC particles accelerates the nozzle erosion for over 150%
comparing to 22 GPa hardness Al2O3 particles (Figure 22). In case of SiC particles, less energy was absorbed by
particle deformation itself and more energy directed to nozzle wall fragmentation. For tungsten carbide nozzle,
Hashish [185] highlighted direct correlation of wear rate with abrasive/nozzle diameters ratio. However, for steel
nozzle, the exit wear rate was stabilizing independently. These opposite trends were explained through the
relatively low erosion resistance of steel but threshold erosion conditions in tungsten carbide.

Figure 22. Influence of abrasive size and hardness on B4C nozzle mass loss [198]

Nanduri et al. [196] provided an empirical model for WC/Co nozzle within AWJM, where nozzle weight
loss rate WN (g/min/mm) is proportional to nozzle pressure P (MPa), abrasive flow rate ma (g/s) and orifice
diameter do (mm) but relates inversely to nozzle length L (mm) and diameter in (mm):
0.38 𝑚0.7
𝑃0.9 𝑑𝑜 𝑎
𝑊𝑁 = (8.07𝐸 – 4) 0.5 𝐿0.8 6-1
𝑑𝑛
Volume removed by a single particle was determined as:
𝐻 7.12
𝑉 ∝ (𝑣 𝑠𝑖𝑛 𝛼)1.9 𝑑 3.28 ( 𝑝) 𝐾𝑡−1.3 6-2
𝐻𝑡
where 𝛼 is the impact angle (around 10°), 𝑣 , 𝑑 and 𝐻𝑝 are the particle velocity, diameter and hardness,
respectively, 𝐻𝑡 and 𝐾𝑡 - target material hardness and fracture toughness. To estimate total volume loss rate, VN,
for other nozzle materials, general dependence was proposed:
𝑚𝑎
𝑉𝑁 = 𝑓𝑁 𝑉 6-3
𝑚𝑝

where 𝑚𝑝 is the mass of a single particle and 𝑓𝑁 is the coefficient, which reflects the effects of jet velocity
distribution, particles fragmentation, secondary impact, etc., that are unknown.

7. Conclusions and prospects

AJM is a progressive manufacturing method with a growing role in the satisfaction of recent and
oncoming industrial demands. With that, future investigations on technology enhancements are required. The
trend of AJM developments is a shift from the macro to micro scale. Further reduction of machining spot, precise
erosion predictability and process controlling are current challenges in AJM.
A variety of AJM methods and developments have been analyzed. Submerged, intermittent and multi-jet
conditions were found to be beneficial for environmental purposes, deep patterning and large area machining,
correspondingly. AAJM is highly competitive in fabrication of surface micro-texture for tribological purposes.
The minimum width of channel achieved by masked micro-blasting is 10 µm [101]. Nevertheless, a feature size
less than 5 μm was suggested for further improvements in surface frictional behaviour [127]. Therefore, the
increase of surface micro-patterning resolution presents an interest in several industries. Liquid-based abrasive jet
polishing has lower MRR than AAJM but provides better surface roughness. With an eye to combine both
advantages, an attempt to build a bridge between air and water-based abrasive jet systems was presented as
CAAJM. Further studies into the understanding of cavitation phenomenon during air-water AJM have the
potential to increase the MRR with high-quality surface preservation. Blasting at cryogenic temperatures gives
the possibility to erode soft materials in a brittle way, eliminating the particles embedment and improving the
surface topography. Future investigations might be directed on the reduction of thermal stress and energy
consumptions of the cooling system. A variety of other strategies of cryogenic cooling during machining [199]
may be tested. Application of high temperatures allows an opposite shift of material removal mechanism. Hard-
to-cut materials can be eroded within semi-ductile mechanism with higher MRR and less energy consumption
using thermally enhanced AJM. Other sources of local heating, for instance laser and plasma, were not studied
but present a powerful prospect for thermally enhanced AJM. It can be noticed from the variety of machining
process that temperature is one of the principal parameters in machining. Investigations of thermal effect onto
removal behaviour and the subsurface damaged layer may have a key value in improving the process efficiency.
Material removal mechanism in brittle materials cannot be entirely expounded by the theory of well-
developed cracks. Failure of relatively homogenous materials is well congruent with the lateral crack propagation
model, although, in materials where the metallic grade is more expressed, cracks penetrate through the boundary
phase between metallic grains. With respect to the fact that microstructure is the primary factor in the erosion
behaviour of ceramics, analytical attempts to include this factor in the material removal mechanism are expected.
Some evidence on ductile cutting of brittle materials at normal impact angles were detected [142]. It was explained
by means of ductile/brittle transition and elevated impact temperature. Although, the investigations on the instant
temperature at impingement spot are almost absent. Since the instant temperature is scarcely empirical subject, a
fundamental study on the thermal side of AJM process is essential. Sub-surface deformations and micro-cracking
can be avoided through the elastic removal mode. Reducing the size of abrasive particles down to 100 nm and
less in AWJM system, material removal is realized mainly through the chemical deposition of atoms from the
substrate to particles. Surface roughness close to 0.1 nm can be achieved, however, an increase in removal rate
remains a research challenge.
Concerning the influence of process parameters, it should be noted that there is no a universal algorithm
for process parameters selection. The effect of separate independent factors may vary significantly or even change
to the opposite. However, the dominant effects of key factors were established. With that, some gaps were found
as well. The effect of impact angle onto surface roughness reveals high interest in AFJP [200] but remains almost
unstudied in AAJM. Such as the main purpose of AAJM is surface texturing, the investigations were mainly
targeted onto prediction of machining profile neglecting its surface roughness and integrity. Since micro-
channelling is usually performed with the use of masks, the influence of abrasive size on footprint width was out
of the research focus. Nowadays, micro-nozzles as small as 130 µm have become commercially available and the
majority of surface patterning tasks can be performed without masks. Moreover, an application of micro-nozzles
may be required for the different combination of SOD, pressure, particles size and other parameters, which may
be accompanied by a new physical phenomenon. Therefore, the influence of process parameters during blasting
with micro nozzles should be revised. An incoming challenge associated with a continuous nozzle and mask
miniaturization is an application of fine abrasive particles. Advanced powder feeding solutions for micro-AAJM
are in demand.
Internal nozzle profile is continuously changing along the axis, which has a negative influence on jet
stability. Nozzle wear rate decelerates with an increase in material hardness, toughness and thermal shock
resistance, nozzle length, and diameter, but rises with operational time, pressure, abrasive flow rate, size, and
hardness. Nozzle life can be significantly improved through optimization of the listed parameters, which has not
been undertaken to date. Multifactorial optimization requires a sufficient empirical database, which is absent till
date. The Laval type of nozzle produces more homogeneous particle distribution inside the jet, substantially
increasing its velocity and improves the footprint geometry. The wear character of the Laval nozzle is not
investigated. The functionally graded nozzle shows a meaningful reduction of erosion rate, that makes it a
promising direction for tool life prolongation. Micro-sapphire and diamond nozzles have recently become
commercially available, but no researches are yet published on its performance.

8. Acknowledgement

The work was supported by the Science Foundation Ireland (SFI) under the Grant Number 15/RP/B3208.

References

[1] Charles et al., Paper metering, cutting and reeling, United States Patent, 1933.
[2] B.G. Schwacha, Liquid cutting of hard materials, United States Patent, 1958.
[3] F. Chen, X. Miao, Y. Tang, S. Yin, A review on recent advances in machining methods based on
abrasive jet polishing (AJP), Int. J. Adv. Manuf. Technol. 90 (2017) 785–799. doi:10.1007/s00170-016-
9405-7.
[4] S. Verma, S.K.Moulick, S.K. Mishra, Nozzle Wear Parameter in Water jet machining The Review, Int.
J. Eng. Dev. Res. 2 (2014) 1063–1073. www.ijedr.org.
[5] H. Syazwani, G. Mebrahitom, A. Azmir, A review on nozzle wear in abrasive water jet machining
application, IOP Conf. Ser. Mater. Sci. Eng. 114 (2016) 2–10. doi:10.1088/1757-899X/114/1/012020.
[6] M. Molitoris, J. Pitel, A. Hosovsky, M. Tothova, K. Zidek, A review of research on water jet with slurry
injection, Procedia Eng. 149 (2016) 333–339. doi:10.1016/j.proeng.2016.06.675.
[7] K. Kalpana, O. V. Mythreyi, M. Kanthababu, Review on condition monitoring of Abrasive Water Jet
Machining system, Proc. 2015 Int. Conf. Robot. Autom. Control Embed. Syst. RACE 2015. (2015).
doi:10.1109/RACE.2015.7097254.
[8] M. Wakuda, Y. Yamauchi, S. Kanzaki, Effect of workpiece properties on machinability in abrasive jet
machining of ceramic materials, Precis. Eng. 26 (2002) 193–198. doi:10.1016/S0141-6359(01)00114-3.
[9] Z.Z. Li, J.M. Wang, X.Q. Peng, L.T. Ho, Z.Q. Yin, S.Y. Li, C.F. Cheung, Removal of single point
diamond-turning marks by abrasive jet polishing, Appl. Opt. 50 (2011) 2458.
doi:10.1364/AO.50.002458.
[10] N. Haghbin, J.K. Spelt, M. Papini, Abrasive waterjet micro-machining of channels in metals:
Comparison between machining in air and submerged in water, Int. J. Mach. Tools Manuf. 88 (2015)
108–117. doi:10.1016/j.ijmachtools.2014.09.012.
[11] H.-C. Li, W.-S. Chen, Recovery of silicon carbide from waste silicon slurry by using flotation, Energy
Procedia. 136 (2017) 53–59. doi:10.1016/J.EGYPRO.2017.10.281.
[12] J.-Y. Kim, U.-S. Kim, M.-S. Byeon, W.-K. Kang, K.-T. Hwang, W.-S. Cho, Recovery of cerium from
glass polishing slurry, J. Rare Earths. 29 (2011) 1075–1078. doi:10.1016/S1002-0721(10)60601-1.
[13] H. Nouraei, A. Wodoslawsky, M. Papini, J.K. Spelt, Characteristics of abrasive slurry jet micro-
machining: A comparison with abrasive air jet micro-machining, J. Mater. Process. Technol. 213 (2013)
1711–1724. doi:10.1016/j.jmatprotec.2013.03.024.
[14] T. Matsumura, T. Muramatsu, S. Fueki, Abrasive water jet machining of glass with stagnation effect,
CIRP Ann. - Manuf. Technol. 60 (2011) 355–358. doi:10.1016/j.cirp.2011.03.118.
[15] K. Kowsari, M. Papini, J.K. Spelt, Selective removal of metallic layers from sintered ceramic and
metallic plates using abrasive slurry-jet micro-machining, J. Manuf. Process. 29 (2017) 252–264.
doi:10.1016/j.jmapro.2017.08.005.
[16] C. Narayanan, R. Balz, D.A. Weiss, K.C. Heiniger, Modelling of abrasive particle energy in water jet
machining, J. Mater. Process. Technol. 213 (2013) 2201–2210. doi:10.1016/j.jmatprotec.2013.06.020.
[17] V. Gupta, P.M. Pandey, M.P. Garg, R. Khanna, N.K. Batra, Minimization of Kerf Taper Angle and Kerf
Width Using Taguchi’s Method in Abrasive Water Jet Machining of Marble, Procedia Mater. Sci. 6
(2014) 140–149. doi:10.1016/j.mspro.2014.07.017.
[18] S. Vasanth, T. Muthuramalingam, P. Vinothkumar, T. Geethapriyan, G. Murali, Performance Analysis
of Process Parameters on Machining Titanium (Ti-6Al-4V) Alloy Using Abrasive Water Jet Machining
Process, Procedia CIRP. 46 (2016) 139–142. doi:10.1016/j.procir.2016.04.072.
[19] R. Shukla, D. Singh, Experimentation investigation of abrasive water jet machining parameters using
Taguchi and Evolutionary optimization techniques, Swarm Evol. Comput. 32 (2017) 167–183.
doi:10.1016/j.swevo.2016.07.002.
[20] D.H. Ahmed, J. Naser, R.T. Deam, Particles impact characteristics on cutting surface during the
abrasive water jet machining: Numerical study, J. Mater. Process. Technol. 232 (2016) 116–130.
doi:10.1016/j.jmatprotec.2016.01.032.
[21] A. Hejjaji, R. Zitoune, L. Crouzeix, S. Le Roux, F. Collombet, Surface and machining induced damage
characterization of abrasive water jet milled carbon/epoxy composite specimens and their impact on
tensile behavior, Wear. 376–377 (2017) 1356–1364. doi:10.1016/j.wear.2017.02.024.
[22] S.T. Kumaran, T.J. Ko, M. Uthayakumar, M.M. Islam, Prediction of surface roughness in abrasive water
jet machining of CFRP composites using regression analysis, J. Alloys Compd. 724 (2017) 1037–1045.
doi:10.1016/j.jallcom.2017.07.108.
[23] Z. Cojbasic, D. Petkovic, S. Shamshirband, C.W. Tong, S. Ch, P. Jankovic, N. Ducic, J. Baralic, Surface
roughness prediction by extreme learning machine constructed with abrasive water jet, Precis. Eng. 43
(2016) 86–92. doi:10.1016/j.precisioneng.2015.06.013.
[24] S. Lathabai, D.C. Pender, Microstructural influence in slurry erosion of ceramics, Wear. 189 (1995)
122–135. doi:10.1016/0043-1648(95)06679-9.
[25] A. V. Levy, G. Hickey, Liquid-solid particle slurry erosion of steels, Wear. 117 (1987) 129–146.
doi:10.1016/0043-1648(87)90251-1.
[26] W. Kordonski, A. Shorey, Magnetorheological (MR) jet finishing technology, J. Intell. Mater. Syst.
Struct. 18 (2007) 1127–1130. doi:10.1177/1045389X07083139.
[27] S. Jha, V.K. Jain, Design and development of the magnetorheological abrasive flow finishing (MRAFF)
process, Int. J. Mach. Tools Manuf. 44 (2004) 1019–1029. doi:10.1016/j.ijmachtools.2004.03.007.
[28] M. Tricard, W.I. Kordonski, A.B. Shorey, Magnetorheological jet finishing of conformal, freeform and
steep concave optics, CIRP Ann. - Manuf. Technol. 55 (2006) 309–312. doi:10.1016/S0007-
8506(07)60423-5.
[29] A.B. Shorey, W.I. Kordonski, S.R. Gorodkin, S.D. Jacobs, R.F. Gans, K.M. Kwong, C.H. Farny, Design
and testing of a new magnetorheometer, Rev. Sci. Instrum. 70 (1999) 4200–4206.
doi:10.1063/1.1150052.
[30] W.K. and S. Gorodkin, Material removal in magnetorheological finishing of optics, QED Technol. Int.
(2011) 1–39.
[31] W. Kordonski, A.B. Shorey, A. Sekeres, New magnetically assisted finishing method: material removal
with magnetorheological fluid jet, Proc. SPIE. 5180 (2004) 107. doi:10.1117/12.506280.
[32] W.I. Kordonski, A.B. Shorey, M. Tricard, Magnetorheological (MR) jet finishing technilogy, Proc.
IMECE04, ASME Int. Mech. Eng. Congr. (2004) 1–8.
[33] W. Kordonski, Jet-induced high-precision finishing of challenging optics, Proc. SPIE. 5869 (2005)
586909-586909–8. doi:10.1117/12.617306.
[34] N. Yuvaraj, M.P. Kumar, Cutting of aluminium alloy with abrasive water jet and cryogenic assisted
abrasive water jet: A comparative study of the surface integrity approach, Wear. 362–363 (2016) 18–32.
doi:10.1016/j.wear.2016.05.008.
[35] H. Getu, Cryogenically assisted abrasive jet micromachining of polymers, J. MICROMECHANICS
MICROENGINEERING. 115010 (2008) 8. doi:10.1088/0960-1317/18/11/115010.
[36] A.G. Gradeen, J.K. Spelt, M. Papini, Cryogenic abrasive jet machining of polydimethylsiloxane at
different temperatures, Wear. 274–275 (2012) 335–344. doi:10.1016/j.wear.2011.09.013.
[37] A.G. Gradeen, M. Papini, J.K. Spelt, The effect of temperature on the cryogenic abrasive jet micro-
machining of polytetrafluoroethylene, high carbon steel and polydimethylsiloxane, Wear. 317 (2014)
170–178. doi:10.1016/j.wear.2014.06.002.
[38] D. Patel, P. Tandon, Experimental investigations of thermally enhanced abrasive water jet machining of
hard-to-machine metals, CIRP J. Manuf. Sci. Technol. 10 (2015) 92–101.
doi:10.1016/j.cirpj.2015.04.002.
[39] H. Fang, P. Guo, J. Yu, Surface roughness and material removal in fluid jet polishing, Appl. Opt. 45
(2006) 4012. doi:10.1364/AO.45.004012.
[40] A. Beaucamp, Y. Namba, Super-smooth finishing of diamond turned hard X-ray molding dies by
combined fluid jet and bonnet polishing, CIRP Ann. - Manuf. Technol. 62 (2013) 315–318.
doi:10.1016/j.cirp.2013.03.010.
[41] A. Beaucamp, Y. Namba, W. Messelink, D. Walker, P. Charlton, R. Freeman, Surface integrity of fluid
jet polished tungsten carbide, Procedia CIRP. 13 (2014) 377–381. doi:10.1016/j.procir.2014.04.064.
[42] Z.-R. Yu, C.-H. Kuo, C.-C. Chen, W.-Y. Hsu, D.P. Tsai, Study of air-driving fluid jet polishing, Opt.
Manuf. Test. 8126 (2011) 812611. doi:10.1117/12.892140.
[43] V. Masa, P. Kuba, Efficient use of compressed air for dry ice blasting, J. Clean. Prod. 111 (2016) 76–84.
doi:10.1016/j.jclepro.2015.07.053.
[44] M. Jerman, H. Orbanić, A. Lebar, I. Sabotin, P. Dreśar, J. Valentinčič, Measuring the water temperature
changes in Ice abrasive water jet prototype, Procedia Eng. 149 (2016) 163–168.
doi:10.1016/j.proeng.2016.06.651.
[45] B. Karpuschewski, T. Emmer, K. Schmidt, M. Petzel, Cryogenic wet-ice blasting - Process conditions
and possibilities, CIRP Ann. - Manuf. Technol. 62 (2013) 319–322. doi:10.1016/j.cirp.2013.03.102.
[46] J.A. McGeough, Cutting of Food Products by Ice-particles in a Water-jet, Procedia CIRP. 42 (2016)
863–865. doi:10.1016/j.procir.2016.03.009.
[47] A.K. Witte, M. Bobal, R. David, B. Blättler, D. Schoder, P. Rossmanith, Investigation of the potential of
dry ice blasting for cleaning and disinfection in the food production environment, LWT - Food Sci.
Technol. 75 (2017) 735–741. doi:10.1016/j.lwt.2016.10.024.
[48] L. Zhang, T. Kuriyagawa, Y. Yasutomi, J. Zhao, Investigation into micro abrasive intermittent jet
machining, Int. J. Mach. Tools Manuf. 45 (2005) 873–879. doi:10.1016/j.ijmachtools.2004.11.003.
[49] T. Kuriyagawa, T. Sakuyama, K. Syoji, H. Onodera, A New Device of Abrasive Jet Machining and
Application to Abrasive Jet Printer, Key Eng. Mater. 196 (2001) 103–110.
doi:10.4028/www.scientific.net/KEM.196.103.
[50] A. Radvanská, T. Ergić, Ž. Ivandić, S. Hloch, J. Valicek, J. Mullerova, Technical possibilities of noise
reduction in material cutting by abrasive water-jet, Strojarstvo. 51 (2009) 347–354.
[51] Y. Murakami, H. Ishii, Erosion due to premixed abrasive water jet under submerged condition, Chem.
Pharm. Bull. 29 (1981) 711–719. doi:10.1248/cpb.37.3229.
[52] N. Haghbin, J.K. Spelt, M. Papini, Abrasive waterjet micro-machining of channels in metals: Model to
predict high aspect-ratio channel profiles for submerged and unsubmerged machining, J. Mater. Process.
Technol. 222 (2015) 399–409. doi:10.1016/j.jmatprotec.2015.03.026.
[53] D. Krewski, R. a Yokel, E. Nieboer, D. Borchelt, J. Cohen, S. Kacew, J. Lindsay, A.M. Mahfouz, V.
Rondeau, Human Health Risk Assessment For Aluminium, Aluminium Oxide, and Aluminium
Hydroxide, 2009. doi:10.1080/10937400701597766.HUMAN.
[54] R.H.M. Jafar, V. Hadavi, J.K. Spelt, M. Papini, Dust reduction in abrasive jet micro-machining using
liquid films, Powder Technol. 301 (2016) 1270–1274. doi:10.1016/j.powtec.2016.08.002.
[55] A.A.A. Novikov F. V., Основи струминно-абразивної обробки дрібних деталей / The basis of
abrasive jet machining of small parts, видавництво ХНЕУ ім. С Кузнеця, м. Харків, 2014.
[56] А.А. Андилахай, Абразивная обработка деталей затопленными струями, ПГТУ, Мариуполь,
2006.
[57] Z. Cao, J.L. Walsh, M.G. Kong, Atmospheric plasma jet array in parallel electric and gas flow fields for
three-dimensional surface treatment, Appl. Phys. Lett. 94 (2009) 021501. doi:10.1063/1.3069276.
[58] J.Y. Kim, J. Ballato, S.O. Kim, Intense and energetic atmospheric pressure plasma jet arrays, Plasma
Process. Polym. 9 (2012) 253–260. doi:10.1002/ppap.201100190.
[59] A. Pan, T. Chen, C. Li, X. Hou, Parallel fabrication of silicon concave microlens array by femtosecond
laser irradiation and mixed acid etching, CHINESE Opt. Lett. 14 (2016) 1–5.
doi:10.3788/COL201614.052201.1.
[60] H. Takino, T. Hosaka, Shaping of steel mold surface of lens array by electrical discharge machining
with single rod electrode, Appl. Opt. 53 (2014) 8002. doi:10.1364/AO.53.008002.
[61] H. Takino, T. Hosaka, Shaping of steel mold surface of lens array by electrical discharge machining
with spherical ball electrode, Appl. Opt. 55 (2016) 4967–4973. doi:10.1364/AO.55.004967.
[62] C.J. Wang, C.F. Cheung, L.T. Ho, M.Y. Liu, W.B. Lee, A novel multi-jet polishing process and tool for
high-efficiency polishing, Int. J. Mach. Tools Manuf. 115 (2017) 60–73.
doi:10.1016/j.ijmachtools.2016.12.006.
[63] F.J. Shiou, A. Asmare, Parameters optimization on surface roughness improvement of Zerodur optical
glass using an innovative rotary abrasive fluid multi-jet polishing process, Precis. Eng. 42 (2015) 93–
100. doi:10.1016/j.precisioneng.2015.04.004.
[64] K. Yamauchi, H. Mimura, K. Inagaki, Y. Mori, Figuring with subnanometer-level accuracy by
numerically controlled elastic emission machining, Rev. Sci. Instrum. 73 (2002) 4028.
doi:10.1063/1.1510573.
[65] R. Balasubramaniam, J. Krishnan, N. Ramakrishnan, Empirical study on the generation of an edge
radius in abrasive jet external deburring (AJED), J. Mater. Process. Technol. 99 (2000) 49–53.
doi:10.1016/S0924-0136(99)00350-7.
[66] S. Barriuso, M. Jaafar, J. Chao, A. Asenjo, J.L. Gonzalez-Carrasco, Improvement of the blasting
induced effects on medical 316 LVM stainless steel by short-term thermal treatments, Surf. Coatings
Technol. 258 (2014) 1075–1081. doi:10.1016/j.surfcoat.2014.07.027.
[67] S. Martens, B. Krueger, W. Mack, F. Voelklein, J. Wilde, Low-cost preparation method for exposing IC
surfaces in stacked die packages by micro-abrasive blasting, Microelectron. Reliab. 48 (2008) 1513–
1516. doi:10.1016/j.microrel.2008.06.033.
[68] Y. Dong, W. Liu, H. Zhang, H. Zhang, On-line recycling of abrasives in abrasive water jet cleaning,
Procedia CIRP. 15 (2014) 278–282. doi:10.1016/j.procir.2014.06.045.
[69] P.J. Slikkerveer, P.C.P. Bouten, F.H. in’t Veld, H. Scholten, Erosion and damage by sharp particles,
Wear. 217 (1998) 237–250. doi:10.1016/S0043-1648(98)00187-2.
[70] N.S. Pawar, R.R. Lakhe, R.L. Shrivastava, Validation of Experimental Work by Using Cubic
Polynomial Models For Sea Sand as an Abrasive Material in Silicon Nozzle in Abrasive Jet Machining
Process, Mater. Today Proc. 2 (2015) 1927–1933. doi:10.1016/j.matpr.2015.07.156.
[71] A. Ghobeity, H. Getu, T. Krajac, J.K. Spelt, M. Papini, Process repeatability in abrasive jet micro-
machining, J. Mater. Process. Technol. 190 (2007) 51–60. doi:10.1016/j.jmatprotec.2007.03.111.
[72] H. Orbanic, M. Junkar, Analysis of striation formation mechanism in abrasive water jet cutting, Wear.
265 (2008) 821–830. doi:10.1016/j.wear.2008.01.018.
[73] A. Momber, Blast cleaning technology, 2008. doi:10.1007/978-3-540-73645-5.
[74] C. Barbatti, J. Garcia, R. Pitonak, H. Pinto, A. Kostka, A. Di Prinzio, M.H. Staia, A.R. Pyzalla,
Influence of micro-blasting on the microstructure and residual stresses of CVD k-Al2O3 coatings, Surf.
Coatings Technol. 203 (2009) 3708–3717. doi:10.1016/j.surfcoat.2009.06.021.
[75] I.T. Kim, Fatigue strength improvement of longitudinal fillet welded out-of-plane gusset joints using air
blast cleaning treatment, Int. J. Fatigue. 48 (2013) 289–299. doi:10.1016/j.ijfatigue.2012.11.010.
[76] D. Arola, A.E. Alade, W. Weber, Improving fatigue strength of metals using abrasive waterjet peening,
Mach. Sci. Technol. 10 (2006) 197–218. doi:10.1080/10910340600710105.
[77] S.A. Meguid, G. Shagal, J.C. Stranart, Finite element modelling of shot-peening residual stresses, J.
Mater. Process. Technol. 92–93 (1999) 401–404. doi:10.1016/S0924-0136(99)00153-3.
[78] K.F. Leong, C.K. Chua, G.S. Chua, C.H. Tan, Abrasive jet deburring of jewellery models built by
stereolithography apparatus (SLA), J. Mater. Process. Technol. 83 (1998) 36–47. doi:10.1016/S0924-
0136(98)00041-7.
[79] R. Balasubramaniam, J. Krishnan, N. Ramakrishnan, Investigation of AJM for deburring, J. Mater.
Process. Technol. 79 (1998) 52–58. doi:10.1016/S0924-0136(97)00305-1.
[80] J. Qu, A.J. Shih, R.O. Scattergood, J. Luo, Abrasive micro-blasting to improve surface integrity of
electrical discharge machined WC-Co composite, J. Mater. Process. Technol. 166 (2005) 440–448.
doi:10.1016/j.jmatprotec.2004.09.075.
[81] A. Beaucamp, Y. Namba, R. Freeman, Dynamic multiphase modeling and optimization of fluid jet
polishing process, CIRP Ann. - Manuf. Technol. 61 (2012) 315–318. doi:10.1016/j.cirp.2012.03.073.
[82] E. Belloy, E. Walckiers, A. Sayah, M.A.M. Gijs, Introduction of powder blasting for sensor and
microsystem applications, Sensors Actuators, A Phys. 84 (2000) 330–337. doi:10.1016/S0924-
4247(00)00390-3.
[83] N. Careddu, O. Akkoyun, An investigation on the efficiency of water-jet technology for graffiti
cleaning, J. Cult. Herit. 19 (2016) 426–434. doi:10.1016/j.culher.2015.11.009.
[84] P. Sanmartín, F. Cappitelli, R. Mitchell, Current methods of graffiti removal: A review, Constr. Build.
Mater. 71 (2014) 363–374. doi:10.1016/j.conbuildmat.2014.08.093.
[85] B.R. Szweda, C. Editor, Jetting into the 21st Century : Water-jet cutting of seals and gaskets, Seal.
Technol. (2001) 6–9.
[86] S. den Dunnen, J. Dankelman, G.M.M.J. Kerkhoffs, G.J.M. Tuijthof, How do jet time, pressure and
bone volume fraction influence the drilling depth when waterjet drilling in porcine bone?, J. Mech.
Behav. Biomed. Mater. 62 (2016) 495–503. doi:10.1016/j.jmbbm.2016.05.030.
[87] G. Kraaij, G.J.M. Tuijthof, J. Dankelman, R.G.H.H. Nelissen, E.R. Valstar, Waterjet cutting of
periprosthetic interface tissue in loosened hip prostheses: An in vitro feasibility study, Med. Eng. Phys.
37 (2015) 245–250. doi:10.1016/j.medengphy.2014.12.009.
[88] S. Den Dunnen, G.J.M. Tuijthof, The influence of water jet diameter and bone structural properties on
the efficiency of pure water jet drilling in porcine bone, Mech. Sci. 5 (2014) 53–58. doi:10.5194/ms-5-
53-2014.
[89] A.J.A. Terzis, G. Nowak, O. Rentzsch, H. Arnold, J. Diebold, G. Baretton, A new system for cutting
brain tissue preserving vessels: Water jet cutting, Br. J. Neurosurg. 3 (1989) 361–366.
doi:10.3109/02688698909002816.
[90] J.A. Zukas, I. May, Water jet for dermatological treatment. United States Patent, 2001.
[91] L. Hallmann, P. Ulmer, E. Reusser, C.H.F. Hämmerle, Effect of blasting pressure, abrasive particle size
and grade on phase transformation and morphological change of dental zirconia surface, Surf. Coatings
Technol. 206 (2012) 4293–4302. doi:10.1016/j.surfcoat.2012.04.043.
[92] H. Costa, I. Hutchings, Some innovative surface texturing techniques for tribological purposes, Proc.
Inst. Mech. Eng. Part J J. Eng. Tribol. 229 (2015) 429–448. doi:10.1177/1350650114539936.
[93] P.J. Slikkerveer, P.C.P. Bouten, F.C.M. De Haas, High quality mechanical etching of brittle materials by
powder blasting, Sensors Actuators, A Phys. 85 (2000) 296–303. doi:10.1016/S0924-4247(00)00343-5.
[94] E. Belloy, A. Sayah, M.A.M. Gijs, Powder blasting for three-dimensional microstructuring of glass,
Sensors Actuators, A Phys. 86 (2000) 231–237. doi:10.1016/S0924-4247(00)00447-7.
[95] D.S. Park, M.W. Cho, H. Lee, W.S. Cho, Micro-grooving of glass using micro-abrasive jet machining,
J. Mater. Process. Technol. 146 (2004) 234–240. doi:10.1016/j.jmatprotec.2003.11.013.
[96] K.D. Bouzakis, G. Skordaris, N. Michailidis, A. Asimakopoulos, G. Erkens, Effect on PVD coated
cemented carbide inserts cutting performance of micro-blasting and lapping of their substrates, Surf.
Coatings Technol. 200 (2005) 128–132. doi:10.1016/j.surfcoat.2005.02.119.
[97] K.D. Bouzakis, A. Tsouknidas, G. Skordaris, E. Bouzakis, S. Makrimallakis, S. Gerardis, G.
Katirtzoglou, Optimization of wet or dry microblasting on PVD films by various Al2O3 grain sizes for
improving the coated tools’ cutting performance, Tribol. Ind. 33 (2011) 49–56.
doi:10.1016/j.cirp.2011.03.012.
[98] H. Li, A. Lee, J. Fan, G.H. Yeoh, J. Wang, On DEM-CFD study of the dynamic characteristics of high
speed micro-abrasive air jet, Powder Technol. 267 (2014) 161–179. doi:10.1016/j.powtec.2014.07.018.
[99] A. Jacob, S. Gangopadhyay, A. Satapathy, S. Mantry, B.B. Jha, Influences of micro-blasting as surface
treatment technique on properties and performance of AlTiN coated tools, J. Manuf. Process. 29 (2017)
407–418. doi:10.1016/j.jmapro.2017.08.013.
[100] D.S. Park, T.I. Seo, M.W. Cho, Mechanical etching of micro pockets by powder blasting, Int. J. Adv.
Manuf. Technol. 25 (2005) 1098–1104. doi:10.1007/s00170-003-1941-2.
[101] H. Wensink, Fabrication of microstructures by Powder Blasting, University of Twente, 2002.
http://core.ac.uk/download/pdf/11465378.pdf.
[102] M. Achtsnick, J. Drabbe, A.M. Hoogstrate, B. Karpuschewski, Erosion behaviour and pattern transfer
accuracy of protecting masks for micro-abrasive blasting, J. Mater. Process. Technol. 149 (2004) 43–49.
doi:10.1016/j.jmatprotec.2003.10.037.
[103] M. Achtsnick, a. M. Hoogstrate, B. Karpuschewski, Advances in High Performance Micro Abrasive
Blasting, CIRP Ann. - Manuf. Technol. 54 (2005) 281–284. doi:10.1016/S0007-8506(07)60103-6.
[104] M. Achtsnick, P.F. Geelhoed, A.M. Hoogstrate, B. Karpuschewski, Modelling and evaluation of the
micro abrasive blasting process, Wear. 259 (2005) 84–94. doi:10.1016/j.wear.2005.01.045.
[105] B. Karpuschewski, A.M. Hoogstrate, M. Achtsnick, Simulation and improvement of the micro abrasive
blasting process, CIRP Ann. - Manuf. Technol. 53 (2004) 251–254. doi:10.1016/S0007-8506(07)60691-
X.
[106] M. Achtsnick, High Performance Micro Abrasive Blasting, Ph.D. Thesis, 2005.
[107] H. Getu, A. Ghobeity, J.K. Spelt, M. Papini, Abrasive jet micromachining of polymethylmethacrylate,
Wear. 263 (2007) 1008–1015. doi:10.1016/j.wear.2007.01.063.
[108] A. Ghobeity, T. Krajac, T. Burzynski, M. Papini, J.K. Spelt, Surface evolution models in abrasive jet
micromachining, Wear. 264 (2008) 185–198. doi:10.1016/j.wear.2007.02.020.
[109] H. Getu, A. Ghobeity, J.K. Spelt, M. Papini, Abrasive jet micromachining of acrylic and polycarbonate
polymers at oblique angles of attack, Wear. 265 (2008) 888–901. doi:10.1016/j.wear.2008.01.013.
[110] A. Ghobeity, D. Ciampini, M. Papini, An analytical model of the effect of particle size distribution on
the surface profile evolution in abrasive jet micromachining, J. Mater. Process. Technol. 209 (2009)
6067–6077. doi:10.1016/j.jmatprotec.2009.05.026.
[111] T. Burzynski, M. Papini, Modelling of surface evolution in abrasive jet micro-machining including
particle second strikes: A level set methodology, J. Mater. Process. Technol. 212 (2012) 1177–1190.
doi:10.1016/j.jmatprotec.2012.01.002.
[112] S. Ally, J.K. Spelt, M. Papini, Prediction of machined surface evolution in the abrasive jet micro-
machining of metals, Wear. 292–293 (2012) 89–99. doi:10.1016/j.wear.2012.05.029.
[113] R.H. Mohammad Jafar, J.K. Spelt, M. Papini, Numerical simulation of surface roughness and erosion
rate of abrasive jet micro-machined channels, Wear. 303 (2013) 302–312.
doi:10.1016/j.wear.2013.03.021.
[114] N. Shafiei, H. Getu, A. Sadeghian, M. Papini, Computer simulation of developing abrasive jet machined
profiles including particle interference, J. Mater. Process. Technol. 209 (2009) 4366–4378.
doi:10.1016/j.jmatprotec.2008.11.020.
[115] A. Ghobeity, M. Papini, J.K. Spelt, Abrasive jet micro-machining of planar areas and transitional slopes
in glass using target oscillation, J. Mater. Process. Technol. 209 (2009) 5123–5132.
doi:10.1016/j.jmatprotec.2009.02.012.
[116] K. Abhishek, S.S. Hiremath, S. Karunanidhi, A novel approach to produce holes with high degree of
cylindricity through Micro-Abrasive Jet Machining (μ-AJM), CIRP J. Manuf. Sci. Technol. (2018).
doi:10.1016/J.CIRPJ.2018.02.002.
[117] H. Li, J. Wang, N. Kwok, T. Nguyen, G.H. Yeoh, A study of the micro-hole geometry evolution on
glass by abrasive air-jet micromachining, J. Manuf. Process. 31 (2018) 156–161.
doi:10.1016/j.jmapro.2017.11.013.
[118] A. Buford, T. Goswami, Review of wear mechanisms in hip implants: Paper I - General, Mater. Des. 25
(2004) 385–393. doi:10.1016/j.matdes.2003.11.010.
[119] S. Kurtz, K. Ong, E. Lau, F. Mowat, M. Halpern, Projections of primary and revision hip and knee
arthroplasty in the United States from 2005 to 2030, J. Bone Jt. Surg. - Ser. A. 89 (2007) 780–785.
doi:10.2106/JBJS.F.00222.
[120] L. Kuncicka, R. Kocich, T.C. Lowe, Advances in metals and alloys for joint replacement, Prog. Mater.
Sci. 88 (2017) 232–280. doi:10.1016/j.pmatsci.2017.04.002.
[121] J. Check, K.S.K. Karuppiah, S. Sundararajan, Comparison of the effect of surface roughness on the
micro/nanotribological behavior of ultra-high-molecular-weight polyethylene (UHMWPE) in air and
bovine serum solution, J. Biomed. Mater. Res. - Part A. 74 (2005) 687–695. doi:10.1002/jbm.a.30355.
[122] A. Wang, V.K. Polineni, C. Stark, J.H. Dumbleton, Effect of Femoral Head Surface Roughness on the
Wear of Ultrahigh Molecular Weight Polyethylene Acetabular Cups, J. Arthroplasty. 13 (1998) 615–
620.
[123] S. Santavirta, Y.. Konttinen, R. Lappalainen, a Anttila, S.. Goodman, M. Lind, L. Smith, M. Takagi, E.
Gómez-Barrena, L. Nordsletten, J.-W. Xu, Materials in total joint replacement, Curr. Orthop. 12 (1998)
51–57. doi:10.1016/S0268-0890(98)90008-1.
[124] D. Dowson, S. Taheri, N.C. Wallbridge, The role of counterface imperfections in the wear of
polyethylene, Wear. 119 (1987) 277–293. doi:10.1016/0043-1648(87)90036-6.
[125] J. Fisher, P. Firkins, E.A. Reeves, J.L. Hailey, G.H. Isaac, The influence of scratches to metallic
counterfaces on the wear of ultra-high molecular weight polyethylene, Proc. Inst. Mech. Eng. Part H J.
Eng. Med. 209 (1995) 263–264. doi:10.1243/PIME_PROC_1995_209_353_02.
[126] L. Caravia, D. Dowson, J. Fisher, B. Jobbins, Proceedings of the Institution of Mechanical Engineers ,
Part H : Journal of Engineering in Medicine The influence of bone and bone cement debris on
counterface roughness in sliding wear tests of ultra-high molecular weight polyethylene on stainless
steel, Proc Instn Mech Engrs. 204 (1990) 65–70. doi:10.1243/PIME.
[127] H. Sawano, S. Warisawa, S. Ishihara, Study on long life of artificial joints by investigating optimal
sliding surface geometry for improvement in wear resistance, Precis. Eng. 33 (2009) 492–498.
doi:10.1016/j.precisioneng.2009.01.005.
[128] S. Wos, W. Koszela, P. Pawlus, The effect of both surfaces textured on improvement of tribological
properties of sliding elements, Tribol. Int. 113 (2017) 182–188. doi:10.1016/j.triboint.2016.10.044.
[129] M.H. Cho, S. Park, Micro CNC surface texturing on polyoxymethylene (POM) and its tribological
performance in lubricated sliding, Tribol. Int. 44 (2011) 859–867. doi:10.1016/j.triboint.2011.03.001.
[130] M. Hasan, J. Zhao, Z. Jiang, A review of modern advancements in micro drilling techniques, J. Manuf.
Process. 29 (2017) 343–375. doi:10.1016/J.JMAPRO.2017.08.006.
[131] S. Agarwal, P.V. Rao, Experimental investigation of surface/subsurface damage formation and material
removal mechanisms in SiC grinding, Int. J. Mach. Tools Manuf. 48 (2008) 698–710.
doi:10.1016/j.ijmachtools.2007.10.013.
[132] Y. Nakanishi, Y. Nakashima, Y. Fujiwara, Y. Komohara, M. Takeya, H. Miura, H. Higaki, Influence of
surface profile of Co-28Cr-6Mo alloy on wear behaviour of ultra-high molecular weight polyethylene
used in artificial joint, Tribol. Int. 118 (2018) 538–546. doi:10.1016/j.triboint.2017.06.030.
[133] A. Borjali, J. Langhorn, K. Monson, B. Raeymaekers, Using a patterned microtexture to reduce
polyethylene wear in metal-on-polyethylene prosthetic bearing couples, Wear. 392–393 (2017) 77–83.
doi:10.1016/j.wear.2017.09.014.
[134] D. Choudhury, R. Walker, T. Roy, S. Paul, R. Mootanah, Performance of honed surface profiles to
artificial hip joints: An experimental investigation, Int. J. Precis. Eng. Manuf. 14 (2013) 1847–1853.
doi:10.1007/s12541-013-0247-z.
[135] T. Matsuno, Reduction of polyethylene wear by concave dimples on the frictional surface in artificial
hip joints, J. Arthroplasty. 15 (2000) 332–338.
[136] M. Buijs, Erosion of Glass as Modeled by indentation Theory, J. Am. Ceram. Soc. 77 (1994) 1676–
1678. doi:10.1111/j.1151-2916.1994.tb09777.x.
[137] R.H.M. Jafar, Erosion and Roughness Modeling in Abrasive Jet Micro-machining of Brittle Materials
Erosion and Roughness Modeling in Abrasive Jet, 2013.
[138] D. Jianxin, F. Yihua, D. Zeliang, S. Peiwei, Wear behavior of ceramic nozzles in sand blasting
treatments, J. Eur. Ceram. Soc. 23 (2003) 323–329. doi:10.1016/S0955-2219(02)00183-8.
[139] J. Deng, Wear behaviors of ceramic nozzles with laminated structure at their entry, Wear. 266 (2009)
30–36. doi:10.1016/j.wear.2008.05.012.
[140] D.B. Marshall, B.R. Lawn, A.G. Evans, Elastic/plastic indentation damage in ceramics: the lateral crack
dystem, J. Am. Ceram. Soc. 65 (1982) 561–566.
[141] B.R. Lawn, A. Evans, D.B. Marshall, Elastic/plastic indentation damage in ceramics: the median/radial
crack system, J. Am. Ceram. Soc. 63 (1980) 574–581. doi:10.1111/j.1151-2916.1982.tb10782.x.
[142] Y. Zhang, Y.B. Cheng, S. Lathabai, Erosion of alumina ceramics by air- and water-suspended garnet
particles, Wear. 240 (2000) 40–51. doi:10.1016/S0043-1648(00)00335-5.
[143] M. Wakuda, Y. Yamauchi, S. Kanzaki, Material response to particle impact during abrasive jet
machining of alumina ceramics, J. Mater. Process. Technol. 132 (2003) 177–183. doi:10.1016/S0924-
0136(02)00848-8.
[144] I. Finnie, Some observations on the erosion of ductile metals, Wear. 19 (1972) 81–90.
doi:10.1016/0043-1648(72)90444-9.
[145] I.M. Hutchings, A model for the erosion of metals by spherical particles at normal incidence, Wear. 326
(1980) 858. doi:10.1111/mec/14011.
[146] B.F. Levin, K.S. Vecchio, J.N. DuPont, A.R. Marder, Modeling solid-particle erosion of ductile alloys,
Metall. Mater. Trans. A. 30 (1999) 1763–1774. doi:10.1007/s11661-999-0175-9.
[147] M. Papini, J.K. Spelt, The plowing erosion of organic coatings by spherical particles, Wear. 222 (1998)
38–48. doi:10.1016/S0043-1648(98)00274-9.
[148] D. Ciampini, J.K. Spelt, M. Papini, Simulation of interference effects in particle streams following
impact with a flat surface Part I. Theory and analysis, Wear. 254 (2003) 237–249. doi:10.1016/S0043-
1648(03)00017-6.
[149] R.R.R. Ellermaa, Erosion prediction of pure metals and carbon steels, Wear. 164 (1993) 1114–1122.
[150] H. Getu, J.K. Spelt, M. Papini, Conditions leading to the embedding of angular and spherical particles
during the solid particle erosion of polymers, Wear. 292–293 (2012) 159–168.
doi:10.1016/j.wear.2012.05.017.
[151] K. Kowsari, J. Schwartzentruber, J.K. Spelt, M. Papini, Erosive smoothing of abrasive slurry-jet micro-
machined channels in glass, PMMA, and sintered ceramics: Experiments and roughness model, Precis.
Eng. 49 (2017) 332–343. doi:10.1016/j.precisioneng.2017.03.003.
[152] K.I. H. Maeda, N. Egami, C. Kagaya, N. Inoue, H. Takeshita, Analysis of particle velocity and
temperature distribution of struck surface in fine particle peening, J. Jpn. Soc. Mech. Eng. (in Japanese).
67 (2001) 306–312. doi:10.1248/cpb.37.3229.
[153] M.A. Tetsuya Senda, Kazuyoshi Arai, Sand erosion of alumina ceramics at elevated temperatures, J.
Ceram. Soc. Japan. 109 (2001) 254–259. doi:10.2109/jcersj.109.1267_254.
[154] H. Wensink, M.C. Elwenspoek, A closer look at the ductile-brittle transition in solid particle erosion,
Wear. 253 (2002) 1035–1043. doi:10.1016/S0043-1648(02)00223-5.
[155] Z. Zhang, W. Liu, Z. Song, Particle size and surfactant effects on chemical mechanical polishing of
glass using silica-based slurry, Appl. Opt. 49 (2010) 5480. doi:10.1364/AO.49.005480.
[156] W. Peng, C. Guan, S. Li, Material removal mode affected by the particle size in fluid jet polishing.,
Appl. Opt. 52 (2013) 7927–33. doi:10.1364/AO.52.007927.
[157] D.M. Small, W.Y. Sanchez, M.J. Hickey, G.C. Gobe, Multiphoton fluorescence microscopy of the live
kidney in health and disease Multiphoton fluorescence microscopy of the live kidney in health and
disease, J. Biomed. Opt. 19 (2014) 020901. doi:10.1117/1.
[158] W. qiang Peng, C. liang Guan, S. yi Li, Defect-free surface of quartz glass polished in elastic mode by
chemical impact reaction, J. Cent. South Univ. 21 (2014) 4438–4444. doi:10.1007/s11771-014-2446-x.
[159] W. Peng, C. Guan, S. Li, Ultrasmooth surface polishing based on the hydrodynamic effect, Appl. Opt.
52 (2013) 6411–6416. doi:10.1364/AO.52.006411.
[160] X. Xu, J. Luo, D. Guo, Nanoparticle-wall collision in a laminar cylindrical liquid jet, J. Colloid Interface
Sci. 359 (2011) 334–338. doi:10.1016/j.jcis.2011.04.001.
[161] Z. Zhang, L. Yu, W. Liu, Z. Song, Surface modification of ceria nanoparticles and their chemical
mechanical polishing behavior on glass substrate, Appl. Surf. Sci. 256 (2010) 3856–3861.
doi:10.1016/j.apsusc.2010.01.040.
[162] K. Yamauchi, K. Hirose, H. Goto, K. Sugiyama, K. Inagaki, K. Yamamura, Y. Sano, Y. Mori, First-
principles simulations of removal process in EEM (Elastic Emission Machining), Comput. Mater. Sci.
14 (1999) 232–235. doi:10.1016/S0927-0256(98)00112-8.
[163] L.M. Cook, Chemical processes in glass polishing, J. Non. Cryst. Solids. 120 (1990) 152–171.
doi:10.1016/0022-3093(90)90200-6.
[164] M. Hashish, A. South, Optimization Factors in Abrasive- Waterjet Machining, J. Eng. Ind. 1 (1991).
[165] N. Su, L. Yue, Y. Liao, W. Liu, H. Zhang, X. Li, H. Wang, J. Shen, The effect of various sandblasting
conditions on surface changes of dental zirconia and shear bond strength between zirconia core and
indirect composite resin, J. Adv. Prosthodont. 7 (2015) 214. doi:10.4047/jap.2015.7.3.214.
[166] N. Menga, R. Di Mundo, G. Carbone, Soft blasting of fluorinated polymers: The easy way to
superhydrophobicity, Mater. Des. 121 (2017) 414–420. doi:10.1016/j.matdes.2017.02.074.
[167] H.Z. Li, J. Wang, J.M. Fan, Analysis and modelling of particle velocities in micro-abrasive air jet, Int. J.
Mach. Tools Manuf. 49 (2009) 850–858. doi:10.1016/j.ijmachtools.2009.05.012.
[168] N. Rajaratnarn, C. Albers, Water distribution in very high velocity water jets in air, J. Hydraul. Eng. 124
(1998) 647–650.
[169] J.M. Fan, H.Z. Li, J. Wang, C.Y. Wang, A study of the flow characteristics in micro-abrasive jets, Exp.
Therm. Fluid Sci. 35 (2011) 1097–1106. doi:10.1016/j.expthermflusci.2011.03.004.
[170] N. Rajaratnam, S.A.H. Rizvi, P.M. Steffler, P.R. Smy, An experimental study of very high velocity
circular water jets in air, J. Hydraul. Res. 32 (1994) 461–470. doi:10.1080/00221689409498746.
[171] K. Kowsari, A. Nouhi, V. Hadavi, J.K. Spelt, M. Papini, Prediction of the erosive footprint in the
abrasive jet micro-machining of flat and curved glass, Tribol. Int. 106 (2017) 101–108.
doi:10.1016/j.triboint.2016.10.038.
[172] M.R. Sookhak Lari, M. Teti, M. Papini, Inverse methods to gradient etch three-dimensional features
with prescribed topographies using abrasive jet micro-machining: Part II - Verification with micro-
machining experiments, Precis. Eng. 45 (2016) 262–271. doi:10.1016/j.precisioneng.2016.03.003.
[173] A. Nouhi, K. Kowsari, J.K. Spelt, M. Papini, Abrasive jet machining of channels on highly-curved glass
and PMMA surfaces, Wear. 356–357 (2016) 30–39. doi:10.1016/j.wear.2016.03.006.
[174] K. Kowsari, M.R. Sookhaklari, H. Nouraei, M. Papini, J.K. Spelt, Hybrid erosive jet micro-milling of
sintered ceramic wafers with and without copper-filled through-holes, J. Mater. Process. Technol. 230
(2016) 198–210. doi:10.1016/j.jmatprotec.2015.11.027.
[175] D. Ciampini, J.K. Spelt, M. Papini, Simulation of interference effects in particle streams following
impact with a flat surface. Part II. Parametric study and implications for erosion testing and blast
cleaning, Wear. 254 (2003) 250–264. doi:10.1016/S0043-1648(03)00016-4.
[176] Q. Li, V. Rudolph, B. Weigl, A. Earl, Interparticle van der Waals force in powder flowability and
compactibility, Int. J. Pharm. 280 (2004) 77–93. doi:10.1016/j.ijpharm.2004.05.001.
[177] F. Podczeck, J.M. Newton, M.B. James, Influence of relative humidity of storage air on the adhesion
and autoadhesion of micronized particles to particulate and compacted powder surfaces, J. Colloid
Interface Sci. 187 (1997) 484–491. doi:10.1006/jcis.1996.4684.
[178] H. Getu, J.K. Spelt, M. Papini, Thermal analysis of cryogenically assisted abrasive jet micromachining
of PDMS, Int. J. Mach. Tools Manuf. 51 (2011) 721–730. doi:10.1016/j.ijmachtools.2011.05.003.
[179] V. Hadavi, B. Michaelsen, M. Papini, Measurements and modeling of instantaneous particle orientation
within abrasive air jets and implications for particle embedding, Wear. 336–337 (2015) 9–20.
doi:10.1016/j.wear.2015.04.016.
[180] A. Nouhi, M.R. Sookhak Lari, J.K. Spelt, M. Papini, Implementation of a shadow mask for direct
writing in abrasive jet micro-machining, J. Mater. Process. Technol. 223 (2015) 232–239.
doi:10.1016/j.jmatprotec.2015.04.007.
[181] A.J. Gant, M.G. Gee, Structure-property relationships in liquid jet erosion of tungsten carbide
hardmetals, Int. J. Refract. Met. Hard Mater. 27 (2009) 332–343. doi:10.1016/j.ijrmhm.2008.11.013.
[182] W. Burwash, W. Finlay, E. Matida, Deposition of particles by a confined impinging jet onto a flat
surface at Re = 10^4, Aerosol Sci. Technol. 40 (2006) 147–156. doi:10.1080/02786820500494551.
[183] J.M. Fan, C.Y. Wang, J. Wang, Modelling the erosion rate in micro abrasive air jet machining of
glasses, Wear. 266 (2009) 968–974. doi:10.1016/j.wear.2008.12.019.
[184] Y.I. Oka, H. Ohnogi, T. Hosokawa, M. Matsumura, The impact angle dependence of erosion damage
caused by solid particle impact, Wear. 203–204 (1997) 573–579. doi:10.1016/S0043-1648(96)07430-3.
[185] M. Hashish, Observations of Wear of Abrasive-Waterjet Nozzle Materials, J. Tribol. 116 (1994) 439.
doi:10.1115/1.2928861.
[186] D. Jianxin, Erosion wear of boron carbide ceramic nozzles by abrasive air-jets, Mater. Sci. Eng. A. 408
(2005) 227–233. doi:10.1016/j.msea.2005.07.029.
[187] D. Jianxin, W. Fengfang, Z. Jinlong, Wear mechanisms of gradient ceramic nozzles in abrasive air-jet
machining, Int. J. Mach. Tools Manuf. 47 (2007) 2031–2039. doi:10.1016/j.ijmachtools.2007.01.011.
[188] Z. Ding, J. Deng, J. Li, Wear surface studies on coal water slurry nozzles in industrial boilers, Mater.
Des. 28 (2007) 1531–1538. doi:10.1016/j.matdes.2006.02.020.
[189] J. Deng, Z. Ding, H. Zhou, Y. Tan, Performance and wear characteristics of ceramic, cemented carbide,
and metal nozzles used in coal-water-slurry boilers, Int. J. Refract. Met. Hard Mater. 27 (2009) 919–
926. doi:10.1016/j.ijrmhm.2009.05.007.
[190] A. Beaucamp, T. Katsuura, Z. Kawara, A novel ultrasonic cavitation assisted fluid jet polishing system,
CIRP Ann. - Manuf. Technol. 66 (2017) 301–304. doi:10.1016/j.cirp.2017.04.083.
[191] A. Sou, S. Hosokawa, A. Tomiyama, Effects of cavitation in a nozzle on liquid jet atomization, Int. J.
Heat Mass Transf. 50 (2007) 3575–3582. doi:10.1016/j.ijheatmasstransfer.2006.12.033.
[192] D. Jianxin, L. Lili, Z. Jinlong, S. Junlong, Erosion wear of laminated ceramic nozzles, Int. J. Refract.
Met. Hard Mater. 25 (2007) 263–270. doi:10.1016/j.ijrmhm.2006.06.005.
[193] D. Jianxin, L. Lili, D. Mingwei, Gradient structures in ceramic nozzles for improved erosion wear
resistance, Ceram. Int. 33 (2007) 1255–1261. doi:10.1016/j.ceramint.2006.03.034.
[194] J. Deng, D. Yun, H. Zhou, Y. Tan, Layered structures in ceramic nozzles for improved erosion wear
resistance in industrial coal-water-slurry boilers, Ceram. Int. 36 (2010) 299–306.
doi:10.1016/j.ceramint.2009.09.003.
[195] J. Deng, L. Liu, M. Ding, Effect of residual stresses on the erosion wear of laminated ceramic nozzles,
Mater. Charact. 59 (2008) 1–8. doi:10.1016/j.matchar.2006.10.009.
[196] M. Nanduri, D.G. Taggart, T.J. Kim, The effects of system and geometric parameters on abrasive water
jet nozzle wear, Int. J. Mach. Tools Manuf. 42 (2002) 615–623. doi:10.1016/S0890-6955(01)00147-X.
[197] F.P.S. M. Nanduri, D.G. Taggart, T.J. Kim, C. Haney, Effect of the inlet taper angle on AWJ nozzle
wear, Proc. 9th Am. Water Jet Conf. 233–235 (1999) 134–150. doi:10.1016/S0043-1648(99)00230-6.
[198] J. Deng, X. Zhang, P. Niu, L. Liu, J. Wang, Wear of ceramic nozzles by dry sand blasting, Tribol. Int.
39 (2006) 274–280. doi:10.1016/j.triboint.2004.07.026.
[199] Y. Yildiz, M. Nalbant, A review of cryogenic cooling in machining processes, Int. J. Mach. Tools
Manuf. 48 (2008) 947–964. doi:10.1016/j.ijmachtools.2008.01.008.
[200] C.J. Wang, C.F. Cheung, M.Y. Liu, Numerical modeling and experimentation of three dimensional
material removal characteristics in fluid jet polishing, Int. J. Mech. Sci. 133 (2017) 568–577.
doi:10.1016/j.ijmecsci.2017.09.018.

View publication stats

You might also like