Inventions: Modern Grinding Technology and Systems

Download as pdf or txt
Download as pdf or txt
You are on page 1of 144

inventions

Modern Grinding
Technology
and Systems
Edited by
W. Brian Rowe
Printed Edition of the Special Issue Published in Inventions

www.mdpi.com/journal/inventions
Modern Grinding Technology
and Systems
Modern Grinding Technology
and Systems

Special Issue Editor


W. Brian Rowe

MDPI • Basel • Beijing • Wuhan • Barcelona • Belgrade


Special Issue Editor
W. Brian Rowe
Liverpool John Moores University
UK

Editorial Office
MDPI
St. Alban-Anlage 66
4052 Basel, Switzerland

This is a reprint of articles from the Special Issue published online in the open access journal Inventions
(ISSN 2411-5134) from 2018 to 2019 (available at: http://www.mdpi.com/journal/inventions/
special issues/MGTS)

For citation purposes, cite each article independently as indicated on the article page online and as
indicated below:

LastName, A.A.; LastName, B.B.; LastName, C.C. Article Title. Journal Name Year, Article Number,
Page Range.

ISBN 978-3-03842-937-1 (Pbk)


ISBN 978-3-03842-938-8 (PDF)


c 2019 by the authors. Articles in this book are Open Access and distributed under the Creative
Commons Attribution (CC BY) license, which allows users to download, copy and build upon
published articles, as long as the author and publisher are properly credited, which ensures maximum
dissemination and a wider impact of our publications.
The book as a whole is distributed by MDPI under the terms and conditions of the Creative Commons
license CC BY-NC-ND.
Contents

About the Special Issue Editor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

Preface to ”Modern Grinding Technology and Systems” . . . . . . . . . . . . . . . . . . . . . . . ix

W. Brian Rowe
Towards High Productivity in Precision Grinding
Reprinted from: Inventions 2018, 3, 24, doi:10.3390/inventions3020024 . . . . . . . . . . . . . . . . 1

Eckart Uhlmann and Joachim Bruckhoff


Modeling and Analysis of Contact Conditions during NC-Form Grinding of Cutting Edges
Reprinted from: Inventions 2017, 2, 13, doi:10.3390/inventions2030013 . . . . . . . . . . . . . . . . 17

Sebastian Barth and Fritz Klocke


Influence of the Grinding Wheel Topography on the Thermo-Mechanical Stress Collective
in Grinding
Reprinted from: Inventions 2017, 2, 34, doi:10.3390/inventions2040034 . . . . . . . . . . . . . . . . 25

Georg Schnurbusch, Ekkard Brinksmeier and Oltmann Riemer


Influence of Cutting Speed on Subsurface Damage Morphology and Distribution in Ground
Fused Silica
Reprinted from: Inventions 2017, 2, 15, doi:10.3390/inventions2030015 . . . . . . . . . . . . . . . . 41

Michal Kuffa, Fredy Kuster and Konrad Wegener


Stochastic Kinematic Process Model with an Implemented Wear Model for High Feed
Dry Grinding
Reprinted from: Inventions 2017, 2, 31, doi:10.3390/inventions2040031 . . . . . . . . . . . . . . . . 53

Mikdam Jamal and Michael N. Morgan


Design Process Control for Improved Surface Finish of Metal Additive Manufactured Parts of
Complex Build Geometry
Reprinted from: Inventions 2017, 2, 36, doi:10.3390/inventions2040036 . . . . . . . . . . . . . . . . 69

Mark J. Jackson and Martin J. Toward


Direct Photonic Fusion of Vitrified Bonding Materials
Reprinted from: Inventions 2017, 2, 19, doi:10.3390/inventions2030019 . . . . . . . . . . . . . . . . 87

Philip Geilert, Carsten Heinzel and André Wagner


Grinding Fluid Jet Characteristics and Their Effect on a Gear Profile Grinding Process
Reprinted from: Inventions 2017, 2, 27, doi:10.3390/inventions2040027 . . . . . . . . . . . . . . . . 95

Peter A. Arrabiyeh, Martin Bohley, Felix Ströer, Benjamin Kirsch, Jörg Seewig and
Jan C. Aurich
Experimental Analysis for the Use of Sodium Dodecyl Sulfate as a Soluble Metal Cutting Fluid
for Micromachining with Electroless-Plated Micropencil Grinding Tools
Reprinted from: Inventions 2017, 2, 29, doi:10.3390/inventions2040029 . . . . . . . . . . . . . . . . 106

Fukuo Hashimoto
Model Development for Optimum Setup Conditions that Satisfy Three Stability Criteria of
Centerless Grinding Systems
Reprinted from: Inventions 2017, 2, 26, doi:10.3390/inventions2040026 . . . . . . . . . . . . . . . . 118

v
About the Special Issue Editor
W. Brian Rowe, Professor. W. Brian Rowe is mainly known for research related to grinding and
grinding machines, including fluid film-bearing technology. His early career centered in the motor
and machine tool industries, leading the research and development of new grinding machines. He
was awarded a Ph.D. and later a Doctor of Science from Manchester University. He became the
Head of Mechanical and Production Engineering at Liverpool Polytechnic, later becoming
Assistant Rector of Liverpool John Moores University. Retiring from this post, he became an
emeritus professor of the University and Director of the Advanced Manufacturing Technology
Research Laboratory, a post he served for more than 10 years. He supervised many doctoral
students, to whom he is eternally grateful for their dedication and enthusiasm in working with
industrial companies to solve their grinding and process control problems.

vii
Preface to ”Modern Grinding Technology
and Systems”
Grinding is a precision material removal process employed to improve the shape, size, and
surface texture of a wide variety of products. The value of the grinding process has been increasingly
recognized because it can achieve high removal rates and very high orders of precision, in many cases
surpassing all other manufacturing processes.
Because grinding constitutes the final machining process controlling the quality and surface
integrity of manufactured parts, it is often described as a ’strategic’ process. It can be considered as
the most important stage of manufacturing, being key to achieving high productivity, high accuracy,
and high quality.
Many countries around the world have recognized the strategic importance of modern grinding
technology and set up specialist research centers to ensure that rapid advances in this field can
be made.
This book reviews these issues in Chapter 1, which includes references to wider sources
of information on specialist aspects. In this chapter, it is shown how engineering research and
engineering ingenuity has led to rapid advances in the productivity and quality of grinding processes.
In some cases, slow lapping and polishing processes for very hard materials have been replaced by
material removal processes conducted at much higher speeds. The succeeding chapters, written by
internationally recognized researchers, illustrate results from investigations of factors that must and
should be controlled to achieve optimum grinding performance. The breadth of the subject matter is
very wide, ranging across materials analysis, abrasives, wheel designs, machine kinematics, grinding
fluids, control systems, and the application of artificial intelligence.
The book will be of interest to industrialists and researchers in the field of manufacturing
technology.

W. Brian Rowe
Special Issue Editor

ix
inventions
Review
Towards High Productivity in Precision Grinding
W. Brian Rowe
General Engineering Research Institute, Liverpool John Moores University, Liverpool L3 3AF, UK;
[email protected]; Tel.: +44-01548-842667

Received: 18 March 2018; Accepted: 10 April 2018; Published: 12 April 2018

Abstract: Over the last century, substantial advances have been made, based on improved
understanding of the requirements of grinding processes, machines, control systems, materials,
abrasives, wheel preparation, coolants, lubricants, and coolant delivery. This paper reviews a
selection of areas in which the application of scientific principles and engineering ingenuity has
led to the development of new grinding processes, abrasives, tools, machines, and systems. Topics
feature a selection of areas where relationships between scientific principles and new techniques
are yielding improved productivity and better quality. These examples point towards further
advances that can fruitfully be pursued. Applications in modern grinding technology range from
high-precision kinematics for grinding very large lenses and reflectors through to medium size
grinding machine processes and further down to grinding very small components used in micro
electro-mechanical systems (MEMS) devices. The importance of material issues is emphasized for the
range of conventional engineering steels, through to aerospace materials, ceramics, and composites.
It is suggested that future advances in productivity will include the wider application of artificial
intelligence and robotics to improve precision, process efficiency, and features required to integrate
grinding processes into wider manufacturing systems.

Keywords: grinding; processes; wheels; machines; systems; control; removal rates; precision; sensors;
micro-grinding; coolant; lubrication; coolant delivery

1. What Is the Potential for Future Innovation


In a narrow context, grinding is the removal of material using a high speed abrasive. In a wider
context abrasive technology includes material removal using low speed abrasive. Innovations that
allow high speed processes to replace low speed processes achieve higher removal rates as explained in
more detail by Rowe [1]. For example, the introduction of new grinding techniques such as electrolytic
in-process dressing using metal-bonded wheels make it possible to mirror-grind hard silicon wafers
with greatly increased removal rates [2].
Examples of low and high speed abrasive processes are illustrated in Figure 1. Higher tool speeds
allow much higher removal rates and have therefore been one of the main drivers of technological
advance. However, this single predominant trend has required many other advances in machine tool
design, grinding wheel design, and system control.
Innovations in modern grinding technology cover an immense range of scientific and industrial
disciplines spanning from the interdisciplinary science of tribology and surface interactions to control
systems and artificial intelligence [1]. The subject fascinates engineers trying to reduce costs and
improve manufacturing productivity within the factory environment and scientists seeking to explain
complex interactions between an abrasive and a workpiece. The drivers of innovation include demands
for better abrasive tools, higher precision, higher removal rates, introduction of new work materials
such as aerospace alloys and hard ceramics, and demands for better surface integrity.

Inventions 2018, 3, 24; doi:10.3390/inventions3020024 1 www.mdpi.com/journal/inventions


Inventions 2018, 3, 24

Figure 1. Increased tool speeds has led to greatly increased material removal rates. (a) Grinding-High
MRR; (b) Honing-Low MRR; (c) Lapping-LowMRR; (d) Polishing-Low MRR

In recent decades, manufacturing engineers are also driven by additional needs such as to operate
in a healthier environment, to eliminate toxic waste, to increase manufacturing flexibility, and to
employ intelligent control of processes using modern digital technology. The potential for innovation
is almost limitless depending ultimately on the engineer’s imagination. For example, lead times
between processes are reduced by integrating a grinding module into a multi-process and multi-tool
system. Manufacturers can also envisage employing a robotic grinding tool together with an automated
optical inspection system to remove product defects within an automated machining system. Looking
back over the previous century, it is startlingly obvious that the range and scope of technological
development has been accelerating, giving great hope for the future.
As manufacturing technology advances, there is increasing awareness that grinding is a “strategic
process” in that it is a critical process for the achievement of the best quality of parts and the lowest
cost [1,3]. This awareness has led to the formation of specialist grinding research laboratories within a
number of universities, particularly in Europe, USA, Japan, and China.
Advances are not always made by a single change of a single element of a grinding system but
usually by a combination of changes affecting several elements. The main parameters the engineer has
to manage are shown in Figure 2, Chen and Morgan [4]. Controlled variables are those inputs that are
open to change either by an engineer, a human operator or by a computer controller. Uncontrolled
variables are basically set values not usually continuously controlled, although potentially open to
change. As the process operates, there is continuous variation in forces, temperatures, vibration,
and grinding power each of which critically affects the output quality of the machined parts and costs.
Parts to be produced vary from very large to very small and from conventional soft engineering
steels to very hard steels, alloys, and ceramics. The diverse range of products required to be machined
means there is a large range of grinding machines having diverse characteristics. It is not proposed to
discuss differences in modern machines and applications in this paper, as it would require too much
space for a single paper. However, a schematic of a special-purpose grinding machine is shown in
Figure 3 as an example of innovative designs produced to achieve high productivity in the grinding of
high precision mirrors and lenses [5]. Whereas most machines are based on an open “C” or “U” shaped
structure for convenience of part-loading, a new generation of high precision machines employ a box
structure to limit machine deflections. With high-precision drives and control systems, the grinding
accuracy achieved was better than 1 μm form accuracy on large 1 m sized parts.

2
Inventions 2018, 3, 24

Figure 2. Important elements of a grinding system. Based on Chen & Morgan [4].

Figure 3. Schematic of Box™ ultra-precision grinder. Shore et al [5].

This paper reviews aspects of science and application leading to improved productivity and
quality in grinding processes. Symbols and nomenclature are listed in Appendix A.

2. Wheel Speeds, Material Properties, and Process Kinematics


It was recognized more than 100 years ago that almost every aspect of process behaviour including
energy efficiency and material removal rate depends on maximum uncut chip thickness as proposed
by Alden [6] and Guest [7]. In this respect, grinding is very little different from micro-milling, although
grinding is performed at much high surface speeds of the cutting edges which √ affects the cutting
mechanics. In simple terms, hmax = 2Lvw ae /vs lc where grit spacing L = 3/Cbcu for a sharp

triangular grit and L = 3/4/Cbcu for a blunt rectangular grit as shown by Rowe ([1], pp. 319).
Increasing chip thickness increases force on the grit. Chip thickness is seen to depend on speeds, feeds,
and depth of cut; also on grit spacing and on grit wear as affected by workpiece, coolant and abrasive
material properties. The major difference between grinding and micro-milling is the randomness of
the grit locations in the wheel surface which means the maximum chip thickness is highly variable
in practice. Variation in maximum chip thickness means that not all abrasive grains cut at optimum
depth. Many will merely rub on the workpiece wasting energy.
Force on the abrasive grit increases with removal rate and reduces with increasing wheel
speed and contact length. However, grinding is also affected by rubbing and ploughing processes.
Figure 4 illustrates three different regimes of physical contact that take place with increasing grain
penetration [8].

3
Inventions 2018, 3, 24

Figure 4. Changes in abrasive regime with increasing grain penetration.

Hahn [8] showed that an abrasive grit rubbing on a ductile material with low normal force
increases or reduces surface roughness without removing significant material. This abrasive action is
typical of blunt grits. With further increase in normal force, a grit plastically deforms the surface causing
side pile-up of material known as ploughing. With yet further increase in force or with a sharper grit,
chips are formed and material is removed from the surface. Advances in grinding technology are often
associated with changes to reduce wasteful rubbing and ploughing action. Innovations aim to increase
the proportion of chip removal compared to ploughing and rubbing so that material is removed with
minimal energy.
The transition points between these three regimes depend on various factors notably abrasive
sharpness, material properties and lubrication. Processes for polishing and lapping traditionally take
place predominantly in the rubbing regime, whereas precision grinding takes place in all three regimes.
High energy deep grinding and cut-off operations take place predominantly in the material removal
regime by chip formation for ductile materials or by crack formation for brittle materials.
It should be emphasized that energy and material removal in grinding is strongly dependent
on the nature of the workpiece material. Hard materials such as cemented carbides [9] and hard
ceramics [10] are brittle and difficult to grind. Such materials tend to exhibit a dominance of crack
formation when machined but also require a very hard abrasive such as diamond. However, in many
cases, it has been shown possible to grind brittle materials in the ductile mode by employing extremely
small chip thickness, thus reducing the risks of weakened structures, a process pioneered by Ohmori
and Nakagawa [2].
Not all brittle materials are difficult to grind. Some materials, such as grey cast iron, are easy to
grind as shown by low energy required compared with grinding hard steels.
Since many parameters influence abrasive contact conditions, there is enormous scope for
achievement of higher process efficiency and for achievement of better surface quality. Process
efficiency depends primarily on the energy required to remove material. Removal rate per unit width:
Q = ae vw . Grinding energy per unit unit volume of material removed is known as specific energy:
e = Ft vs /ae vw . Figure 5 shows typical relationships between Q and e.
Figure 5a based on Comley et al. [11], shows results for high efficiency deep grinding of camshaft
webs at very high removal rates using a sharp electro-plated superabrasive cBN wheel. Figure 5b for
surface grinding is based on Malkin [12], and shows results for low speed shallow grinding at 30 m/s
using conventional wheels to grind the material AISI 1095HR. Comparing Figure 5a,b, it is seen that
huge increases in removal rate of up to 100 times have been achieved in high efficiency deep grinding
(HEDG) using high-speed superabrasive wheels together with effective total system design.
Both sets of results show that increasing removal rate by increasing ae and vw , reduces specific
energy asymptotically. The asymptotic energy level is typically of the order of 10 J/mm3 but can even
be as low as 5 J/mm3 , depending on the work material and abrasive [13].

4
Inventions 2018, 3, 24

Figure 5. Typical results show specific energy reduces with increasing removal rates. (a) Based on
Comley et al. [11]; (b) Based on Malkin [12].

Malkin [12] concluded that the energy carried away with the chips must be lower than the
asymptotic value and that the energy corresponds to the enthalpy for the maximum chip temperature,
ech = ρcTch The energy required before steel chips start to become molten, is typically about 6 J/mm3 ,
approaching a value close to the melting energy of the work material. As melting temperatures are
approached, the energy required to shear the material, is reduced. This process and a predominance of
chip formation energy over rubbing and sliding energy provides a partial explanation for a low energy
asymptote at high removal rates.
Unfortunately, a limit is found to removal rates, when the work material starts to flow into the
grinding wheel surface and the abrasive action breaks down. Some materials such as inconel 718,
maintain hardness at high temperatures, which makes the material difficult to grind. High forces
on the abrasive grits cause rapid wear and high temperatures that increase the tendency for wheel
loading. It may be necessary to reduce chip thickness by increasing wheel speed and also to ensure
effective lubrication and cooling [14].

3. Process Limits to Removal Rate in Grinding


Increases in removal rate tend to be limited by the need to maintain or improve product quality
as specified for accuracy, surface integrity, and surface roughness. Relationships between quality
parameters and removal rate depends on many factors such as kinematics, material properties, abrasive
properties, machine stiffness, cooling, and lubrication as indicated in Figure 2. Over the years,
innovation has allowed removal rates to be increased often with improved quality as a result of
increased knowledge, improved understanding of process behaviour and ingenuity employed to
expand the process boundaries. Figure 6 shows a limit chart derived for centreless grinding [15].

Figure 6. Typical process limit chart for centreless plunge grinding AISI 1055 steel 50 mm dia. × 65 mm
long at wheel speeds ranging from 30 m/s to 60 m/s. Based on Rowe et al. [15].

5
Inventions 2018, 3, 24

Typical process limits are shown in Figure 6, for infeed rate and work speed where removal
rate increases with infeed rate v f according to Q = πdw v f /2 in centreless grinding or Q = πdw v f
for centre grinding. The safe operating range for the particular grinding wheel, machine design,
and work material is enclosed by a thermal damage boundary, the machine power limit, and the
chatter limit. This chart is typical for many grinding processes although the nature of the limits may
vary. For example, the power limit shown may in many cases be replaced by a maximum surface
roughness limit or possibly by a size control limit. The striking feature of the figure is that the maximum
infeed rate is greatly increased by increasing the wheel speed from 30 m/s to 60 m/s. This was partly
explained by the reduction in chip thickness with increasing wheel speed and corresponding reduction
in grinding forces, but was also partly explained by the characteristics of the main spindle speed
controller. Much higher grinding speeds, more than 150 m/s have since been made possible by use of
superabrasive wheels [16]. It is seen in Figure 6 that low work speeds and high infeed rates lead to
burn. It is also found that chatter vibration becomes more of a problem at high work speeds. It can
therefore be seen for a particular machine set-up, there is an optimum work speed where a maximum
feed rate can be achieved.

4. Process Control for High Productivity and Quality, Use of Sensors, and Process Models
Process control may involve introducing variations in process input parameters to correct for
quality errors such as inaccuracy, surface integrity problems due to thermal damage, excessive
roughness, and excessive vibration. With repeated batch manufacture of a product, it is possible
to apply process control changes between parts produced. Sometimes, with larger workpieces or
longer grinding cycles, it may be necessary to apply corrections within a part cycle.
Unfortunately, it is not usually possible to employ data directly from charts such as Figure 6 for
process control. This is because the process boundaries shift with time depending on various factors
such as grinding wheel wear, workpiece hardness, redressing kinematics, and dressing tool wear.
An alternative is to employ process sensors and process models to monitor process changes and to
predict necessary changes to bring a process back into a safe operating region.
A major advance in grinding process control in the 20th Century was the introduction of precision
diameter gauging linked to CNC control. Typical results for precision grinding with diameter gauging
are shown in Figure 7. Typical results for a constant feedrate system are shown in Figure 7a. Employing
a 60 s cycle time, it was possible to hold size for successive parts within approximately 1 μm. However,
with an adaptive control system, it was possible to predict a new target feed position and to vary the
spark-out dwell period for each part [17]. A similar size range was achieved but mean cycle time was
reduced from 60 s to 37 s, Figure 7b. With adaptive control, removal rate was substantially improved
without sacrificing accuracy.

Figure 7. Size variations using diameter gauging with position control: (a) For constant feed-rate and
(b) for adaptive power and dwell. Based on Liverton & Rowe [17].

Using adaptive control, it is necessary to use sensors that can detect whether the system is being
maintained within an acceptable operating region while increasing the control variable towards the
process limit. Various sensors have been used, in addition to the standard position sensors employed
for CNC, the most common being a power sensor which is reasonably inexpensive and easy to
incorporate into a system without requiring a dynamometer fixture. Other sensors, including size

6
Inventions 2018, 3, 24

sensors, force sensors, acoustic emission sensors, and Barkhausen noise sensors may be employed to
check for safe operation. To overcome many problems, such as burn or chatter, it may be necessary to
reduce work speed or to reduce infeed rate or possibly to redress the grinding wheel. Some decisions
require an intelligent input either from a human operator or possibly from AI software incorporated
into the control system. For process control, it is often necessary to employ a process model to relate
a measured variable to the required control variable. For example, to adaptively control spark-out
dwell period, as in Figure 7b, the best way was to determine system time constant from the measured
power [18].
For prevention of thermal damage in industrial processes, a power sensor is usually employed,
either in conjuction with a thermal model or without a thermal model [1,3]. Of course, there are other
decisions that affect the onset of thermal damage such as redressing a worn wheel or even replacing
the grinding wheel with a more efficient wheel.
The onset of thermal damage correlates with an excessive grinding power for the particular
grinding conditions or an excessive force level. In Figure 8, using a force sensor, feed rate is
increased in small steps up to the limit for thermal damage, Steffan et al. [19]. The application
of post-process Barkhausen noise measurement for correlation with temperature modeling is described
by Sridharan et al. [20]. This latter paper introduces the concept of modelling and correlating more
than one measured variable for improved reliability of prediction. The use of multiple sensors will no
doubt be a trend for the future.

Figure 8. Adaptive control of feed-rate avoiding thermal damage. Based on Steffan et al. [19].

5. Temperature Rise Modeling


Thermal models may be used as a basis for process control, Rowe [21]. Temperature rise is
critical for workpiece surface hardness, surface roughness, material phase changes, and onset of tensile
residual stress. The Rowe thermal model has been validated for a range of grinding processes including
conventional shallow grinding, creep feed grinding, and HEDG [21].
The model can be derived from heat balance of the heat flows in the contact zone. Four heat flows
make up the total heat flow, Werner et al. [22]. The total heat flow
results from the grinding power P dissipated within the grinding contact area lc bw and is defined
as q ≡ P/lc bw . The four heat flows take place to the workpiece, the abrasive grains, the chips and
the grinding fluid. The chip energy qch , see Section 2, is immediately carried away from the contact
area and plays no further part in heat partitioning. The remaining heat q − qch , is shared at the grain
and workpiece rubbing interface according to Hahn [8] and also according to the more refined model
of Rowe et al. [21,23]. The heat flow into the workpiece at the grain contact is given by Rws (q − qch ).
Some heat q f is conducted out again by the fluid within the grinding contact area. The temperature
rise, of the workpiece is therefore given by the Rowe thermal model [1,21], as follows:
 
ΔT = Rws (q − qch )/ hw + h f

7
Inventions 2018, 3, 24

where, Rws has a value between 0 and 1 depending on the grain conductivity, the workpiece thermal
properties, the grit contact area, and the wheel speed. For conventional abrasives when dry grinding
steels, a typical value is 0.85, and for superabrasive cBN, a typical value is 0.46. For diamond
abrasive a typical value is even lower, showing a great advantage of superabrasives for avoiding large
temperature rise.
The terms hw and h f are convection factors for the workpiece and the fluid respectively.
The convection factor for the workpiece is given by classical heat transfer theory and defined as
hw ≡ qw /ΔT given by, 
βw vw
hw = .
C lc
where C ≈ 1 for conventional speeds. For shallow cuts and deep cuts and other values of speed, values
of C are given in Figure 9 and in the literature [1,21]. Temperatures on the finish surface are seen to be
lower than on the grinding contact surface for deep grinding.

Figure 9. C-factors for temperature rise where Peclet Number L = vw lc /4α.

It can be seen that a process model for fluid film convection is highly desirable to form a
link between grinding energy and workpiece temperature rise. This is still a subject of research.
An expression sometimes used to estimate fluid convection assumes the pores of the wheel are filled
with cold fluid throughout the contact arc and cools the workpiece as though it was a “solid-fluid

wheel”. The convection factor from this assumption is given by h f = ( β f /C ). vs /lc . Correlation
with experimental results, Figure 10 shows reasonable agreement in low temperature grinding below
the fluid boiling temperature [24]. It is also found that fluid convection can be much higher than
previously realized. In practice, the convection factor rapidly reduces towards zero if the fluid boils.

Figure 10. Model for water based fluid agrees best for grinding temperature below 100 C. Data
from [24].

8
Inventions 2018, 3, 24

Simplifications, for example, ignoring fluid convection, conduction into the wheel, and chip
energy, over-estimate temperature rise and grossly over-estimate temperature rise in some cases to the
point that steel would apparently reach 10,000 ◦ C or more. A simplified but inaccurate assumption
√ √
that all heat enters the workpiece, gives the misleading expression ΔT = q lc /β w vw .

6. Grinding Fluids and Fluid Delivery


Grinding fluids play several important roles in a liquid-based wet grinding process. As seen in
the previous section, a grinding fluid can play a valuable role in reducing grinding temperature by
providing surface cooling within the grinding contact area. A major early review of friction, cooling,
and lubricaton in grinding lists 160 research papers (Brinksmeier et al. [25]). Other roles are played
by the grinding fluid including keeping the machine cool, keeping the body of the workpiece cool
and flushing away the grinding debris. This latter role is particularly important in high removal rate
grinding and in mirror finish grinding. Grinding swarf can clog a grinding wheel or be pulled into the
grinding contact area causing surface damage [26]. The grinding fluid also serves a physical–chemical
lubrication role in the contact between the work material and the abrasive material which can be
critical for avoidance of excessive grinding wheel wear and rapid onset of wheel loading [27].
There are three primary classes of grinding fluids: (1) fluids which are mixed with water known
as water-based fluids and emulsions; (2) mineral oils; and (3) synthetic oils. Within these groups there
are a very diverse family of oils and chemicals selected according to the type of abrasive and the nature
of the work material.
Water-based coolants have a better cooling property than oils but boil at a much lower temperature.
The much reduced cooling above the boiling temperature of water based coolants is seen in
Figure 9. Water-based coolants also need to be changed frequently before excessive contamination and
bio-degradation take place. Neat oil coolants have better chemical stability and are generally better for
preventing subsequent corrosion of the workpieces.
Coarser grit wheels allow more fluid to circulate into the grinding contact and are generally better
at maintaining the cutting efficiency of the wheel [28]. The supply of fluid into the grinding zone
is impeded however by the boundary layer that surrounds a speeding grinding wheel [29]. The air
barrier can be overcome by providing a fluid jet velocity at a similar speed to the wheel speed and an
ample supply of fluid [30]. However, it is wasteful to supply too much fluid as the grinding surface
is unable to accommodate a large excess in the pores. This can be seen from Figure 11. If the fluid
supply is adequate, the quantity of fluid that actually enters the grinding contact increases with wheel
speed. However, when the wheel speed is too high, the supply of fluid is no longer sufficient so that
the useful flow levels off and will actually start to reduce. The shape and profile of the nozzle affects
the effectiveness of delivery. Round nozzles are generally better than rectangular nozzles but a single
round nozzle cannot cope with wide wheels. In either case, an internal concavely converging feed to
the nozzle is better than a convexly converging nozzle [31].
It is possible to roughly optimize the velocity and the flowrate required to maximize the useful
flow that enters the grinding contact quite simply based on the wheel porosity, taken from the wheel
manufacturer’s data, the wheel speed and the wheel width [32]. Of course, there are times when much
smaller quantities of fluid should be applied, as when finishing a product to achieve a very fine size
dimensional tolerance. The fluid causes a hydrodynamic force that pushes the grinding wheel away
from the ground surface. Under these circumstances, it is necessary to reduce the removal rate and
reduce the flowrate to limit the hydrodynamic force.
There is a cost of fluid cooling both financial and environmental [33]. Fluid supply requires energy
to provide the high velocities and supply rates needed for high removal rate grinding at high wheel
speeds. Unless the jet velocity is sufficient, the fluid jet causes a drag on the grinding wheel because
the fluid has to be accelerated up to grinding wheel speed. The power required taken through the
wheel drive motor can be substantial, particularly when using a shoe nozzle.

9
Inventions 2018, 3, 24

Environmental concerns have led to the introduction of minimum quantity lubrication (MQL)
for grinding [34]. MQL is usually more appropriate for low removal rate grinding since the cooling
achieved is much lower than with liquid coolants. Conventional cooling delivery for grinding may
involve many litres/minute whereas the delivery for MQL is usually in the form of a fine mist of esters
or mineral oils injected into an air stream involving millilitres/min. Good results have been achieved
for some finish grinding operations.

Figure 11. Useful flowrate for a high-porosity grinding wheel. Based on Gviniashvili [30].

7. Grinding Wheels and Abrasives


Conventional grinding wheels are largely based on vitrified bond aluminium oxide abrasives
or on vitrified bond silicon carbide abrasives, although other abrasives are employed for specialist
purposes. For example, for flexible wheels, resin, rubber or polymer bonds are often employed.
Most conventional vitrified grinding wheels run at speeds of 20–45 m/s although with rigorous
attention to safety requirements, vitrified wheels can be obtained for higher speeds. Monolayer
superabrasive grinding wheels, such as diamond and cBN wheels often run at much higher speeds,
in excess of 150 m/s. Although, superabrasives are more expensive, much greater hardness allows
these materials to achieve much longer life when grinding very hard mechanical engineering and
electronic materials.
In more recent decades, there have been substantial developments in abrasive technology and
wheel designs allowing very high speeds to be used as detailed by the late Mike Hitchiner formerly
of Saint Gobain [16]. Examples of conventional silicon carbide and alumina grains are shown in
Figure 12a,b, courtesy of Saint Gobain. Whereas conventional grains tend to be more blocky in shape,
a new generation of more advanced sintered grain abrasives, as shown in Figure 12c,d, have very
large length to width ratio. This allows the grains to present a sharper profile when grinding and
also allows harder grains to be employed to yield longer wheel life. Large length to width ratio of the
grains accommodates more fluid in the pores for porosity and better cooling. The manufacturers tend
to refer to such structures as superabrasive ceramics. The new structures have proven very successful
in high removal rate grinding with harder materials that quickly blunt more conventional abrasives.
Figure 12e,f shows examples of brazed and electro-plated monolayer superabrasive CBN wheel
structures. Monolayer wheels allow very high wheel speeds [35]. Multilayer vitrified CBN wheels are
also available as a replacement for conventional wheels. CBN is not only harder than conventional
abrasives but has excellent heat conduction properties resulting in lower grinding temperatures [1,36].

10
Inventions 2018, 3, 24

High speed vitrified wheels requires extreme care to design for the stresses involved and avoid
wheel failures [37]. This is because the vitrified bond is the source of potential failure whereas
monolayer and multilayer metal-bond wheels provide the capability to run at the highest speeds.

Figure 12. Conventional abrasives, superabrasive SG, and CBN. Courtesy of Saint Gobain Abrasives.
(a) Green silicon carbide; (b) white alumina; (c) sintered SG; (d) extruded SG; (e) Brazed CBN;
(f) EP CBN.

Diamond is harder than CBN, but is generally unsuited to grinding steels because carbon diffuses
from the diamond causing rapid wheel wear. However, diamond due to its hardness has advantages
for very hard materials.
Conventional vitrified bond wheels are generally dressed using diamond tools to correct the
profile accuracy and to sharpen the wheel surface. Dressing time is non-productive so it is an advantage
if wheels can grind for a long time before needing to be redressed [38]. Vitrified CBN wheels may
also be dressed but dressing consumes valuable CBN material. A technique to minimize the material
removed in dressing and to optimize subsequent grinding performance is known as ‘touch dressing’.
Acoustic emission sensors detect the instant of contact between the dressing tool and the wheel, hence
allowing accurate setting of a minimal dressed depth [39].
Monolayer CBN and monolayer diamond wheels are not usually dressed since this would damage
the abrasive layer. Such wheels must therefore be manufactured with great accuracy to ensure the
grains all lay at the same level on the surface. Manufacturers aim to control the evenness and circularity
of the wheel to 1 micron.
Some modern dense materials such as hard ceramics require diamond or CBN grinding because
of their hardness and brittle characteristics. It is often possible to grind such materials and obtain
mirror surface finishes using extremely small depth of cut. The depth of cut has to be small enough
that material removal takes place primarily in a ductile mode. Large depths of cut lead to brittle
mode deformation causing surface cracks and probable failure. Processes developed in recent decades
involve the use of fine-grained CBN or diamond in metal-bond wheels. Dressing such wheels has been
developed using electro-chemical and electrolytic removal processes. For example, a particular process

11
Inventions 2018, 3, 24

known as in-process electrolytic dressing, (ELID) is employed in a grinding process to mirror-grind


silicon wafers [2,40]. Electrolytic dressing exposes the abrasive grains and thus allows a grinding
process to take place. Grinding replaces a lapping process with much improved productivity and
accuracy. Basic elements of an ELID system are shown in Figure 13.

Figure 13. Electrolytic in-process dressing system.

In recent decades, a need has been apparent for grinding wheels that can grind micro-tools
which may have a tip diameter of a few tens of micrometers, Ohmori et al. [41]. Such tools have
been successfully ground using the ELID process. Aurich et al., also describe the machining of
micro-parts [42]. Figure 14 shows relatively large micro-tools produced by Butler-Smith et al. [43].
Figure 14a shows a tool where fine diamond grains are secured on the surface by electro-plating and
Figure 14b shows a tool where the regular pattern of cutting edges was produced by laser ablation of
solid diamond formed by chemical vapour deposition (CVD). The regularity and even spacing of the
cutting edges of the laser ablated tool was said to give improved surface roughness and accuracy.

Figure 14. Micro-grinding wheels: (a) electro-plated diamond wheel; (b) laser-ablated diamond wheel.
Photographs supplied and reproduced by permission of P W Butler-Smith.

The recent rapid increase in parts produced by additive manufacturing has created a need for
finishing processes that can improve the surface roughness of parts having complex geometrical
shapes. An alternative to grinding using a grinding wheel is the use of abrasive mass flow processes
as described by Jamal and Morgan [44].

8. Application of Artificial Intelligence and System Integration


Modern CNC systems contain intelligent software for various functions such as position, velocity,
and acceleration control. AI generally implies decision-making that is, to some extent, judgemental

12
Inventions 2018, 3, 24

based on sensor input. It implies that the machine can make a decision that otherwise may require a
human being. Whereas a human being may become tired, the AI machine will continuously perform
repetitive tasks without human intervention, making adjustments where necessary. AI may be used to
help control the grinding process by reacting to sensors within the grinding machine and measurement
equipment as mentioned in Section 4. AI analyses and translates data from measurements to make
decisions such as to increase or reduce feedrate to avoid thermal damage or to initiate a change point
in a feed cycle as described above or to raise an alarm call for operator intervention. The application of
standard AI tools such as artificial neural networks, genetic algorithms, and case-based reasoning in
grinding systems are further discussed in a review paper [45].
With the passage of time it becomes clear, that the ability of computers to collect, process,
and analyse data is becoming ubiquitous. Data can be collected and analyzed for control and
scheduling of manufacture in a wider sense. For the user and indeed for the machine tool manufacturer
there is the possibility to collect data concerning the need for maintenance of the grinding machine
or the effectiveness of its operation, using the Internet of Things (IoT). A further way in which AI
can help is to assist in selecting appropriate tools or grinding conditions for the initial process set-up.
An example is the selection of an appropriate grinding wheel including wheel grade and structure
using case-based reasoning [46].

9. Conclusions
This paper reviews a selection of developments in which the application of science and ingenuity
has led to new grinding processes, abrasives, tools, machines, and systems. It is shown that improved
technology has yielded higher productivity and better quality across a diverse range of applications
involving very large and very small products. Modern grinding technology has learned to cope with a
diverse range of materials relying on new abrasive tools and various kinematic operating conditions.
It is concluded that this trend will continue and that future advances in productivity will include the
wider application of artificial intelligence and robotics to improve precision, process efficiency, and
features required to integrate grinding processes into wider manufacturing systems and into versatile
machine tools.

Appendix A. Notation
ae Real depth of cut
bcu Uncut grit contact width
bw Width of grinding contact
c Specific heat capacity of work material
C Number of active grits per unit area
de Equivalent wheel diameter
e or u Specific energy to remove unit volume of material
hm Maximum uncut chip thickness
vs Wheel surface speed
k Thermal conductivity
lc Grinding contact length
L Effective tangential grit spacing or
L Peclet number
P Grinding power
q Heat flux given as heat flow per unit contact area
qch Heat flux to chips
qw Heat flux to workpiece within contact area
qs Heat flux to wheel within contact area
qf Heat flux to grinding fluid within contact area
Q Removal rate per unit width

13
Inventions 2018, 3, 24

ΔT Temperature rise
Tch Temperature rise of chips
vf Normal feed or infeed rate
vs Wheel surface speed
vw Workpiece surface speed
α Thermal diffusivity of work material

β Thermal property kρc
ρ Mass density of work material
∅ = lc /de Grinding contact angle

References
1. Rowe, W.B. Modern Grinding Technology, 2nd ed.; Elsevier, William Andrew Imprint: Waltham, MA, USA;
Oxford, UK, 2014.
2. Ohmori, H.; Nakagawa, T. Mirror Surface Grinding of Silicon Wafers with Electrolytic in Process Dressing.
CIRP Ann. 1990, 39, 329–332. [CrossRef]
3. Inasaki, I.; Toenshoff, H.K.; Howes, T.D. Abrasive Machining in the Future. CIRP Ann. 1993, 42, 723–732.
[CrossRef]
4. Chen, X.; Morgan, M.N. Advances in Quality and Productivity in Precision Grinding—A Review of Selected
Research. In Proceedings of the ASME International Manufacturing Science & Engineering Conference,
Symposium on Abrasive Machining, Blacksburg, VA, USA, 27 June–1 July 2016.
5. Shore, P.; Morantz, P.; Luo, X.; Tonnelier, X.; Collins, R.; Roberts, A.; May-Miller, R.; Read, R. Big OptiX
Ultra-Precision Grinding/Measuring System. Proc. SPIE 2005, 5965, 241–248.
6. Alden, G.L. Operation of Grinding Wheels in Machine Grinding. ASME Trans. 1914, 36, 451–460.
7. Guest, J.J. Grinding Machinery; Edward Arnold: London, UK, 1915.
8. Hahn, R.S. On the Nature of the Grinding Process. In Proceedings of the 3rd Machine Tool Design
and Research Conference, Birmingham University, Birmingham, UK, September 1962; Pergamon Press:
Oxford, UK, 1962; pp. 129–154.
9. Wirtz, C.; Mueller, S.; Mattfeld, P.; Klocke, F. A Discussion on Material Removal Mechanisms of Cemented
Carbides. J. Manuf. Sci. Eng. 2017, 139, 121002. [CrossRef]
10. Marinescu, I.D.; Toenshoff, H.K.; Inasaki, I. (Eds.) Handbook of Ceramic Grinding and Polishing;
Noyes Publications: Park Ridge, NJ, USA; William Andrew Publishing: Norwich, NY, USA, 2000.
11. Comley, P.; Stephenson, D.J.; Corbett, J. High Efficiency Deep Grinding and the Effect on Surface Integrity.
Key Eng. Mater. 2004, 257–258, 207–212. [CrossRef]
12. Malkin, S. Grinding Technology; Ellis Horwood: Chichester, UK, 1989.
13. Rowe, W.B.; Jin, T. Temperatures in High Efficiency Deep Grinding. CIRP Ann. 2001, 50, 205–208. [CrossRef]
14. Rowe, W.B.; Ebbrell, S. Process Requirements for Cost-Effective Precision Grinding. CIRP Ann. 2004, 53,
255–258. [CrossRef]
15. Rowe, W.B.; Bell, W.F.; Brough, D. Optimization Studies in High Removal Rate Centreless Grinding.
CIRP Ann. 1986, 35, 235–238. [CrossRef]
16. Marinescu, I.D.; Hitchiner, M.P.; Uhlmann, E.; Rowe, W.B.; Inasaki, I. Handbook of Machining with Grinding
Wheels, 2nd ed.; CRC Press, Taylor and Francis: Boka Raton, FL, USA, 2016.
17. Liverton, J.; Rowe, W.B. Adaptive Control of Cylindrical Grinding—From Development to
Commercialization. In Proceedings of the 5th International Society of Manufacturing Engineers Grinding
Conference, Cincinnati, OH, USA, 26–28 October 1993.
18. Thomas, D.A.; Allanson, D.R.; Moruzzi, J.L.; Rowe, W.B. In-process Identification of Time Constant for the
Control of Grinding. J. Eng. Ind. 1995, 117, 194–201. [CrossRef]
19. Steffan, M.; Haas, F.; Pierer, A.; Jens, G. Adaptive Grinding Process—Prevention of Thermal Damage Using
OPC-UA Technique and In Situ Metrology. J. Manuf. Sci. Eng. 2017, 139, 121008. [CrossRef]
20. Sridharan, U.; Bedekar, V.; Kolarits, F. A Functional Approach to Integrating Temperature Modeling and
Barkhausen Noise Analysis for Prediction of Surface Integrity in Bearing Steels. CIRP Ann. 2017, 66, 333–336.
[CrossRef]
21. Rowe, W.B. Temperatures in Grinding—A Review. J. Manuf. Sci. Eng. 2017, 139, 121001. [CrossRef]

14
Inventions 2018, 3, 24

22. Werner, P.G.; Younis, M.A.; Schlingensiepen, R. Creep-feed—An Effective Method to Reduce Workpiece
Surface Temperatures in High Efficiency Grinding Processes. In Proceedings of the 8th North American
Metalworking Research Conference, Rolla, MO, USA, 18–21 May 1980; Soc. of Manuf. Engineers; pp. 312–319.
23. Rowe, W.B.; Black, S.; Mills, B.; Qi, H. Analysis of Grinding Temperatures by Energy Partitioning. Proc. Inst.
Mech. Eng. Part B 1996, 210, 579–588. [CrossRef]
24. Zhang, L.; Rowe, W.B.; Morgan, M.N. An Improved Fluid Convection Solution in Conventional Grinding.
Proc. Inst. Mech. Eng. Part B 2013, 227, 832–838. [CrossRef]
25. Brinksmeier, E.; Heinzel, C.; Wittmann, M. Friction, Cooling and Lubrication in Grinding. CIRP Ann. 1999,
48, 581–598. [CrossRef]
26. Heinzel, C.; Antsupov, G. Prevention of Wheel Clogging in Creep-feed Grinding by Efficient Tool Cleaning.
CIRP Ann. 2012, 61, 323–326. [CrossRef]
27. Marinescu, I.D.; Rowe, W.B.; Dimitrov, B.; Ohmori, H. Tribology of Abrasive Machining Processes;
William Andrew: Waltham, MA, USA; Oxford, UK, 2013.
28. Engineeer, F.; Guo, C.; Malkin, S. Experimental Measurement of Fluid Flow through the Grinding Zone.
J. Eng. Ind. 1992, 114, 61–66. [CrossRef]
29. Ebbrell, S.; Woolley, N.H.; Tridimas, Y.D.; Allanson, D.R.; Rowe, W.B. The Effects of Cutting Fluid Application
on the Grinding Process. Int. J. Mach. Tools Manuf. 2000, 40, 209–223. [CrossRef]
30. Gviniashvili, V.; Webster, J.; Rowe, W.B. Fluid Flow and Pressure in the Grinding Wheel Workpiece-Interface.
J. Manuf. Sci. Eng. 2005, 127, 201–205. [CrossRef]
31. Rouly, E.; Bauer, R.J.; Warkentin, A. An investigation into the effect of nozzle shape and jet pressure in profile
creepfeed grinding. Proc. Inst. Mech. Eng. Part B. J. Eng. Manuf. 2017, 231, 1116–1130. [CrossRef]
32. Morgan, M.N.; Jackson, A.R.; Baines-Jones, V.; Batako, A.D.; Rowe, W.B. Optimization of Fluid Application
in Grinding. CIRP Ann. 2008, 57, 363–366. [CrossRef]
33. Howes, T.D.; Toenshoff, H.K.; Heuer, W. Environmental Aspects of Grinding Fluids. CIRP Ann. 1991, 40,
623–630. [CrossRef]
34. Tawakoli, T.; Hada, M.J.; Sadeghi, M.H.; Daneshi, A.; Stockart, S.; Rasifard, A. An Experimental Investigation
of the Effects of Workpiece Grinding Parameters in MQL. Int. J. Mach. Tools Manuf. 2009, 49, 924–932.
[CrossRef]
35. Webster, J.; Tricard, M. Innovations in Abrasive Products for Precision Grinding. CIRP Ann. 2004, 53, 597–642.
[CrossRef]
36. Hitchiner, M.P.; McSpadden, S. Evaluation of Factors Controlling CBN Abrasive Selection for Vitrified
Bonded Wheels. Key Eng. Mater. 2004, 257–258, 267–272. [CrossRef]
37. Barlow, N.; Rowe, W.B. Discussion of Stresses in Plain and Reinforced Cylindrical Grinding Wheels. Int. J.
Mach. Tool Des. Res. 1983, 23, 153–160. [CrossRef]
38. Rowe, W.B.; Chen, X. The Identification of Dressing Strategies for Optimal Grinding Performance.
In Proceedings of the Thirtieth International MATADOR Conference, Manchester, UK,
31 March–1 April 1993; pp. 195–202.
39. Rowe, W.B.; Chen, X.; Allanson, D.R. The Coolant Coupling Method Applied to Touch Dressing in High
Frequency Internal Grinding. In Proceedings of the Thirty-Second International Matador Conference,
Manchester, UK, July 1997; UMIST, Macmillan Press: Oxford, UK, 1997; pp. 337–340.
40. Lui, J.H.; Pei, Z.J.; Fisher, G.R. ELID Grinding of Silicon Wafers: A Literature Review. Int. J. Mach. Tools Manuf.
2007, 47, 529–536.
41. Ohmori, H.; Katahira, K.; Narusa, T.; Uehara, Y.; Nakao, A.; Mizutani, M. Microscopic Effects on Fabrication
of Ultra-Fine Micro Tools. CIRP Ann. 2007, 56, 569–572. [CrossRef]
42. Aurich, J.C.; Engmann, J.; Schueler, G.M.; Haberland, R. Micro-Grinding Tool for Manufacture of Complex
Structures in Brittle Materials. CIRP Ann. 2009, 58, 311–314. [CrossRef]
43. Butler-Smith, P.W.; Axinte, D.A.; Daine, M. Solid Diamon Tools: From Innovative Design and Fabriaction to
Preliminary Performance Evaluation in Ti–6Al–4V. Int. J. Mach. Tools Manuf. 2012, 59, 55–64. [CrossRef]
44. Jamal, M.; Morgan, M.N. Design Process Control for Improved Surface Finish of Metal Additive
Manufactured Parts of Complex Build Geometry. Inventions 2017, 2, 36. [CrossRef]

15
Inventions 2018, 3, 24

45. Rowe, W.B.; Li, Y.; Inasaki, I.; Malkin, S. Applications of Artifical Intelligence in Grinding. CIRP Ann. 1994,
43, 521–532. [CrossRef]
46. Li, Y.; Mills, B.; Rowe, W.B. An Intelligent System for Selection of Grinding Wheels. Proc. Inst. Mech. Eng.
Part B J. Eng. Manuf. 1997, 211, 635–641. [CrossRef]

© 2018 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

16
inventions
Article
Modeling and Analysis of Contact Conditions during
NC-Form Grinding of Cutting Edges
Eckart Uhlmann and Joachim Bruckhoff *
Institute for Machine Tools and Factory Management, Technical University Berlin, 10623 Berlin, Germany;
[email protected]
* Correspondence: [email protected]; Tel.: +49-30-314-23473

Received: 31 May 2017; Accepted: 29 June 2017; Published: 5 July 2017

Abstract: Due to increasing demands on cutting tools, cutting edge preparation is of high priority
because of its influence on the tool life. Current cutting edge preparation processes are mostly
limited to generating simple roundings on the cutting edge. Multi-axis high precision form grinding
processes offer great potential to generate defined cutting edge microgeometries. Knowledge about
the relation between grinding strategy and material removal rate can achieve improved work results
with regard to higher precision of shape and dimensional accuracy as well as enhanced cutting edge
quality. Therefore, a kinematic-geometric model was developed in order to analyze the complex
contact conditions during grinding cutting edge microgeometries by using a simulation approach
based on the intersection of geometric bodies. The subsequent grinding tests largely validated the
utilized simulation approach.

Keywords: NC-form grinding; cutting edge; contact conditions

1. Introduction
Increasing demands on cutting tools with regard to quality and economy have led to constant
progress in tool development and tool production. The cutting edge in particular is attracting much
interest as it is an essential tool element for the cutting process. The properties of the cutting edge
influence the process characteristics and the work results in machining with defined cutting edges [1,2].
The properties of the cutting edge can be divided into physical and chemical characteristics as well as
into the geometry and surface topography of the cutting edge. The macro and micro geometrical design
of the cutting edge affects the chip formation and the chip removal [3,4]. Investigations focusing on the
effect of the cutting edge shape on tool wear in turning experiments showed that cutting edges rounded
by brushing have a higher stability and therefore do not tend to have cutting edge breakouts [5]. Further
studies about different methods for cutting edge preparation of micro mills made of tungsten carbide
showed that magnetic abrasive finishing and drag finishing result in favorable wear behavior because
of homogeneous cutting edge profiles [6]. The most common production processes for cutting edge
preparation are magnetic abrasive finishing, abrasive blasting, laser machining, brushing, abrasive
flow machining, vibratory finishing and drag finishing. These production processes usually generate
simple cutting edge roundings. Until now, systematic studies on the influence of different defined
cutting edge microgeometries were not possible, as all common cutting edge preparation processes
are not able to produce complex and exact defined forms. Developing manufacturing processes to
produce defined and complex cutting geometries in the industrial and scientific environment can
increase the performance of cutting processes. Multi-axis high precision form grinding processes offer
great potential to generate defined cutting edge microgeometries.
Due to its positive hardness and toughness characteristics, cemented carbide is often used as a
material in cutting tools [7]. Most cemented carbides consist of a cobalt matrix and the hard material

Inventions 2017, 2, 13; doi:10.3390/inventions2030013 17 www.mdpi.com/journal/inventions


Inventions 2017, 2, 13

tungsten carbide. The cutting material is produced in a sinter process. The grain size of the sintered
powder, as well as the mixing ratio, influences the properties of the material. Due to the high hardness
of tungsten carbide of approximately Vicker Hardness 2000 HV0.05, economical machining of cemented
carbide can be achieved by grinding processes only [8].
NC (Numerical Control)-form grinding processes with mounted points are used for
manufacturing-free forms on difficult-to-machine materials. Examples of these are the machining
of optical glasses and ceramic implants or prostheses for dental and medical applications.
The main challenges in NC-form grinding are the positional accuracy and the stiffness of the
machine axes due to the simultaneous axis movement. Furthermore, the complex tool paths
mostly require NC-programming with CAD/CAM (Computer-aided Design/Computer-aided
Manufacturing)technology. Tool paths are graphically planned on CAD models and converted into
an appropriate NC-Code by post-processors.
Different form grinding strategies lead to different courses of chip cross-sectional areas and thus
different material removal rates. As a result, different process forces occur that could lead to varying
tool deflections and defects of the cutting edge. Knowledge about the relation between grinding
strategy and material removal rate can lead to improved work results in the form of higher preciseness
of shape and dimensional accuracy as well as enhanced cutting edge quality. CAD/CAM-programs do
not provide a proper calculation of chip cross-sectional areas and thus material removal rates. Hence,
it is necessary to develop suitable simulation models.
To enable a reproducible manufacturing of cutting edge geometries, different strategies for form
grinding of cutting edge microgeometries, with regard to material removal rates by using permeation
simulations, are examined. In grinding tests, the spindle power for the different grinding strategies is
compared with the course of the graphs of the material removal rate. The aim of this analysis is to
assess these different form grinding strategies with regard to material removal rates to increase the
process performance and work results.

2. Test Conditions
The program to simulate the permeation of the tool and the workpiece was developed with the
program software MATLAB, The MathWorks Inc., Natick, MA, USA. The form grinding processes
of the cutting edge micro geometries were executed on a five-axis-machining center RXP600DSH of
Röders GmbH, Soltau, Germany. Besides three linear axes (X, Y, Z), this machine is equipped with two
rotatory axes (A, C). Furthermore, the integrated grinding spindle has a maximum power of 8.5 kW
to enable a maximum rotation speed of 60,000 rpm. To dress the grinding tools, a dressing spindle
with a sintered diamond form roller is integrated. To produce the required sharpness of the mounted
points, a sharpening block is used and an automated sharpening process using the through-feed
method was performed. Moreover, the grinding machine enables the determination of the geometry
and the reference of the part with the help of a 3D-probe. The workpieces were cutting inserts made
of cemented carbide of the type SNMA120408. As cooling lubricant, grinding oil of MKU-Chemie
GmbH, Rödermark, Germany, with a viscosity of 7.6 mm2 /s at 40 ◦ C was used. Moreover, metal
bonded diamond mounted points with a grain size of D20 were utilized. Only the use of diamond
tools allows an economic grinding processing of cemented carbides. To reduce tool wear and thus
dimensional deviations on the workpiece, metal bond tools were used. An important advantage of
metal bonds is the shape stability of the tool profiles [9]. In order to gain process forces, conventional
force measurement platforms cannot be satisfactorily used when grinding with 5-axis kinematics
because of relocations of the self weight. Rotating multicomponent dynamometers could be suitable
solutions but the non-contact signal transmission caused by the design principle can be used only
at maximum spindle speeds of approximately 12,000 rpm. Therefore, the performance data of the
grinding spindle was recorded and used to validate the simulation results. To verify the accuracy of
this measurement method, the performance data was compared to force values recorded by a force
measurement platform during a peripheral longitudinal grinding process, as shown in Figure 1.

18
Inventions 2017, 2, 13

Figure 1. Comparison of calculated mechanical power Pc with directly measured electric spindle power P.

It is shown that, with the used process parameters, both force progressions, from depths of cut of
approximately 200 μm, show a good correlation. During manufacturing, cutting edge microgeometries
with much smaller material removal rates than in Figure 1 will be realized. To create equally reliable
statements between spindle power data and process forces, the material removal rate was increased by
offsetting the target geometry in the workpiece, Figure 2.

Figure 2. Offset of nominal geometry to increase material removal rates.

Due to the relatively long tool paths of the mounted point with a constant depth of cut on the rake
and flank face, these constant values of spindle power can be seen as the reference during grinding
of the cutting edge. Therefore, it is possible to identify the actual machining of the cutting edge.
The difference between the spindle power minimum and maximum is shown in the further graphs as
a percentage increase of spindle performance in comparison to the minimum.

3. Kinematic-Geometric Model

3.1. Cutting Edge Microgeometries and Tool Types


In the investigations, two different cutting edge geometries were considered: an ideal radius and
a waterfall radius. Both microgeometries can be characterized using the form-factor method. The form
factor is the relationship between the cutting edge segment on the rake face Sγ and the cutting edge
segment on the flank face Sα . The ideal radius has a form factor κ = 1. Therefore, the cutting edge

19
Inventions 2017, 2, 13

segments on flank and rake faces, as well as the cutting edge radius, have the same value. The waterfall
radius is characterized by a form factor greater than one, in this case κ = 1.5, Figure 3.

Figure 3. Form-factor method for ideal radius and waterfall radius.

For both cutting edge microgeometries, different grinding strategies were examined, as shown
in Figure 4. Grinding strategy I and II differ with regard to the active tool shape. In Strategy I,
the sphere-shaped form of the cylinder ball end mounted point is used to cut orthogonally above
the cutting edge, referred to as the course of the corner radius rε . Strategy II has the same tool
kinematics but the cylindrical part of the mounted point is utilized for the cutting process. In Strategy
III, the ball-shaped part of the mounted point is guided above the cutting edge at an angle of 45◦ ,
referred to as the course of the corner radius. For machining the waterfall geometries with a form
factor κ = 1.5, the grinding strategies I and II match with the strategies of the ideal grinding radius.

Figure 4. Different grinding strategies for NC-form grinding cutting edge microgeometries.

3.2. Modelling of the Permeation


The aim of the developed process simulation is to identify the material removal rates of different
grinding strategies and cutting edge microgeometries for one tool path movement. Thus, knowledge
about forces involved and therefore about tool deflection can be generated to analyze different grinding
strategies. The kinematic-geometrical model simulates the penetration of two geometrical bodies.
The workpiece (cutting insert) is defined as a cylindrical surface and the mounted point is also defined
as a cylindrical surface if the required tool type is cylindrical; if the required tool type is spherical,
an appropriate partial surface is created, Figure 5. The geometric objects are described as point clouds.

20
Inventions 2017, 2, 13

To reduce the computing time, the program minimizes the point clouds to the points of the contacting
areas of both geometric objects.

Figure 5. Schematic representation of the tool and workpiece areas relevant for simulated calculation.

The target geometry of the respective cutting edge microgeometry is precisely positioned as
a numerical approximated curve in the workpiece point cloud. This curve is the reference for the radial
position of the most external point of the tool point cloud. For each calculation step, the tool point
cloud is placed at the following point of the cutting edge curve, from the flank face (angle position
α = 0) to the rake face (angle position α = 90◦ ). Therefore, the rotation of the approximated tool around
the approximated cutting edge enables the determination of the exact section planes. In Figure 6,
the green cycles represent the reduced tool point cloud and the blue area represents the reduced
workpiece point cloud.

Figure 6. Intersection of tool and workpiece point clouds in one position on one tool path.

The red curve depicts the cutting curve of the cutting edge target geometry. The point distance of
both point clouds is 0.1 μm to achieve high calculation accuracy. For each position of the tool point
cloud, the intersection points are projected on a plane. Afterwards, the area within the external points
can be determined which is equivalent to the section plane between the tool and workpiece, Figure 7.

21
Inventions 2017, 2, 13

Figure 7. Generating the cut surface by locating the outer intersections.

Before running the simulation, the boundary conditions have to be defined. These include
geometry data of the cutting edge microgeometry and the tool. Further simulation data is the feed
speed and the resolution of geometric objects influencing the accuracy of the calculations.

4. Results

4.1. Simulation Results


The results of the simulation show the different material removal rate Qw curve progressions for
each grinding strategy, as shown in Figure 8. The lowest maximum material removal rate Qw,max is
achieved by grinding strategy II.

Figure 8. Simulation results for different form grinding strategies.

22
Inventions 2017, 2, 13

The maximum value is located at an angle position α of approximately 20◦ . Therefore, the maximum
material removal rate Qw,max of grinding strategy I was determined at the start angle position of the
mounted point, referred to as the course of the rounded cutting edge radius rβ . It is the highest achieved
value. Following the further angle positions, the curve is regressive and has low values from angle
position α = 16◦ . The course of grinding strategy III is similar but has a higher peak than grinding strategy
II. Therefore, it does not appear advantageous for process forces and consequently for tool deflections.
Hence, further considerations relate to grinding strategies I and II. The courses of the different grinding
strategies I and II for grinding waterfall radii with a form factor κ = 1.5 differ only slightly from the
courses of the equivalent grinding strategies for the ideal radii, Figure 8. With an angle position of α = 70◦ ,
almost all the material is removed. This is caused by the larger curvature at the beginning of the cutting
edge geometry and by the fact that the ratio of the tool radius to the cutting edge radius is approximately
100. This results in relatively flat tool geometry, as shown in Figure 6.

4.2. Experimental Results


To validate the kinematic-geometrical model, grinding tests were conducted. For this purpose,
the spindle power of the different grinding strategies was recorded and compared with the courses of
the material removal rates from the simulation calculations, as shown in Figure 9. Generally, for all test
series, the courses of the spindle power are highly similar to the courses of the simulation calculations.
However, small course deviations can be identified.

Figure 9. Comparison of curve progressions of simulation data and spindle power data.

For both cutting edge geometries, slight deviations in the spindle power courses can be observed
in comparison to the simulation results, Figure 9. The reason for this occurrence may be the inertia of
the tool spindle’s continuous turn regulation in combination with the comparatively short process time
of one tool path of approximately 8 s. Another important factor is the relatively small spindle load
and therefore the less linear behavior between the material removal rate and process forces, as well
as the correct determination of the relevant part of the spindle power course. Nevertheless, analysis
of the spindle power data obtained in the grinding tests confirms the trend of the simulation results.
The validation of the simulation model confirms the benefit of form grinding processes of cutting edge
microgeometries for planning strategies. The kinematic-geometrical model enables the prediction
of maximum material removal rates and subsequently indicates process forces and potential tool
deflections. Furthermore, the influence of the grinding strategy on possible damages to the cutting
edge can be analyzed.

23
Inventions 2017, 2, 13

5. Summary
Within the presented work, a kinematic-geometric model was developed in order to analyze the
complex contact conditions during grinding cutting edge microgeometries. Therefore, a simulation
approach based on the intersection of geometric bodies was used. The grinding tool and workpiece
surface were defined as cylindrical point clouds and the cross section of the cutting edge microgeometry
was determined by numerical approximation. By moving the tool point cloud stepwise along the
coordinates of the cutting edge microgeometry for each position, the interface between the workpiece
point cloud was calculated. Thus, the varying material removal rates could be determined and
compared with the spindle power data. The relatively even curve progressions show the correlation
between the material removal rate Qw and the process forces. Therefore, the geometric-kinematic
model can be used to analyze different form grinding strategies for manufacturing various cutting edge
microgeometries with regard to process forces and consequently tool deflection. With the gathered
information, the work results of form grinding cutting edge microgeometries, regarding the form
accuracy and the chipping of the cutting edge, can be improved.

Acknowledgments: The authors would like to thank the Deutsche Forschungsgemeinschaft (German Research
Foundation) for funding this research within the project DFG UH 100/179-1.
Author Contributions: Both authors were equally involved in creating this paper.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Biermann, D.; Wolf, M.; Aßmuth, R. Cutting edge preparation to enhance the performance of single lip deep
hole drills. Procedia CIRP 2012, 1, 172–177. [CrossRef]
2. Klocke, F.; Kratz, H. Advanced tool edge geometry for high precision hard turning. CIRP Ann. Manuf. Technol.
2005, 54, 47–50. [CrossRef]
3. Denkena, B.; Biermann, D. Cutting edge geometries. CIRP Ann. Manuf. Technol. 2014, 63, 631–653. [CrossRef]
4. Kümmel, J.; Braun, D.; Gibmeier, J.; Schneider, J.; Greiner, C.; Schulze, V.; Wanner, A. Study on micro
texturing of uncoated cemented carbide cutting tools for wear improvement and built-up edge stabilization.
J. Mater. Process. Technol. 2015, 215, 62–70. [CrossRef]
5. Denkena, B.; Grove, T.; Bergmann, B. Eine frage des radius. WB 2016, 149, 58–61.
6. Uhlmann, E.; Oberschmidt, D.; Löwenstein, A.; Polte, M.; Winker, I. Schneidkantenpräparation von
VHM-Mikrofräsern. wt Werkstattstechnik Online 2015. Available online: http://www.werkstattstechnik.de/
wt/article.php?data%5Barticle_id%5D=84726 (accessed on 30 May 2017).
7. Gleim, P. Untersuchungen Zum Bandsägen Mit Diamantbeschichteten Werkzeugen; Kassel University Press GmbH:
Kassel, Germany, 2006.
8. Schedler, W. Hartmetall Für Den Praktiker: Aufbau, Herstellung, Eigenschaften Und Industrielle Anwendungen
Einer Modernen Werkstoffgruppe; VDI-Verlag: Düsseldorf, Germany, 1988; ISBN 3184008037.
9. Denkena, B.; Tönshoff, H.K. Spanen Grundlagen; Springer: Heidelberg, Germany; Dordrecht, The Netherlands;
London, UK; New York, NY, USA, 2011; ISBN 978-364219772-7.

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

24
inventions
Article
Influence of the Grinding Wheel Topography on the
Thermo-Mechanical Stress Collective in Grinding
Sebastian Barth * and Fritz Klocke
Laboratory for Machine Tools and Production Engineering (WZL), RWTH Aachen University,
52074 Aachen, Germany; [email protected]
* Correspondence: [email protected]; Tel.: +49-241-8028183

Received: 16 October 2017; Accepted: 29 November 2017; Published: 4 December 2017

Abstract: The grinding process is used for both high-performance machining and surface finishing of
hardened steel. In addition to the grinding parameters and the grinding fluid supply, the topography
of the grinding wheel mainly determines the grinding process behavior and the grinding process
result. An alteration of the topography by a variation of the volumetric composition of the grinding
wheel, by a variation of the grinding wheel conditioning, or by wear causes a change in the contact
conditions. The state of the art shows a substantial knowledge deficit about the influence of
the volumetric grinding wheel composition and the resulting grinding wheel topography on the
thermo-mechanical stress collective acting on the workpiece external zone. Thus, it is not possible
to make a quantitative statement about the influence of the volumetric grinding wheel composition
on the external zone properties of a component after grinding. Therefore, the aim of the current
research is an empirical-analytical model for the prediction of the thermo-mechanical stress collective
as a function of the grinding wheel topography. For this purpose, a methodology is developed,
which enables the prediction of the topography-dependent thermo-mechanical load in a grinding
process. Therefore, the topography is characterized by means of quantitative parameters and the
main influencing variables on the grinding process behavior are investigated. The findings are used
to analyze the influence of a change in the topography on the grinding temperature and the grinding
force. The obtained results are summarized and are used to explain the thermo-mechanical stress
collective as a function of the grinding wheel topography.

Keywords: grinding; topography; thermo-mechanical stress collective; cutting edge; contact conditions

1. Introduction
The topography of a grinding wheel represents a functional surface. It is part of a tribological system
and causes a stress collective during the engagement with the workpiece material. It determines the
chipping of the workpiece material, and thus, the stresses, such as mechanical loads and temperatures.
To identify the influence of the grinding wheel topography on this thermo-mechanical stress collective,
the topography of the grinding wheel must be described adequately. In principle, the grinding wheel
topography is dependent on the grinding wheel specification [1,2]. Further, it is strongly influenced
by the grinding-wheel conditioning process and geometrical changes during the grinding process due
to wear [3,4]. The characteristics of the topography determine the machining behavior in the grinding
process, particularly, so it must be taken into account when designing grinding processes [5].
In the past, several approaches were developed to characterize the grinding wheel topography
and to investigate its influence on the grinding behavior. Nowicki proposed a method for selecting
topographic characterization parameters for a multi-parameter representation based on low statistical
correlation analysis. Based on 100 machined and ground surfaces, he constructed one cross-correlation
matrix. Nine parameters were compared, including three height parameters, four spacing parameters,

Inventions 2017, 2, 34 25
1 www.mdpi.com/journal/inventions
Inventions 2017, 2, 34

asperity radius, and slope. The three height parameter pairs were found to have correlation coefficients
above 0.9, but for the other 29 pairs, he did not find any correlation [6]. Hecker et al. presented a
three-dimensional methodology to evaluate the static parameters of a topography to calculate the
dynamic parameters. In their model, the individual grain geometries were approximated to a conical or
a spherical shape. They found that primarily the protrusion of the grains determines the dynamic grain
density and the grinding force [7]. For the characterization of the topography, Nguyen and Butler used
the root-mean square parameter Sq to describe its coarseness and the parameters density of summits
Sds and summits curvature Ssc to indicate the abrasive grain density and their sharpness. The results
show a significant correlation between the parameters Sds and Sq, but the results do not show a
correlation between the topography and the investigated grinding force [8]. Uhlmann et al. described
the surface of a grinding wheel to develop a reliable method for Numerical Control (NC)-grinding by
using a single profile of kinematic cutting edges instead of the whole topography. The simplification
of the grinding tool topography proved to be feasible for grinding with abrasive mounted points,
in which rotations are high and feed rates are low. A quantitative description of the topography or the
kinematic cutting edges is not given [5].
According to Hübert et al., the sufficient characterization of the grinding wheel topography is only
possible with the aid of value-supported height, surface, and volume data [9]. Since 2007, DIN ISO 25178
standard [10] has consistently been used to control the acquisition and determination of surface-related
and three-dimensional surface parameters. In order to characterize a surface, it is differed between
surface-related height parameters and functions, including related function parameters.
The height parameters include the mean arithmetic height of a surface Sa, which indicates the
arithmetic mean value of the absolute ordinate values within a definition range, and the mean square
height of a surface Sq, which calculates the root from the average square value of the ordinal values of
a range. The skewness of a surface Ssk, which corresponds to the quotient of the mean third power of
the ordinate values and the third power of Sq, is a measure of the symmetry of the amplitude density
curve. Furthermore, the kurtosis of the surface Sku, the steepness of the height distribution within the
definition area, as well as the maximum peak and maximum sink height Sp and Sv, and the maximum
height Sz (sum of Sp and Sv), are counted to the height characteristics.
Furthermore, the standard specifies the definition of functions for describing the surface texture
and the related parameters. In particular, the area-based material ratio curve of scale-limited surfaces,
also known as the Abbott-Curve, is of interest in the description of the grinding wheel topography.
This function represents the material portion of the scale-bound surface as a function of height, and thus
describes the increase in the proportion of material (Mr) of a surface with the increasing depth of the
roughness profile.
The Abbott-Curve is formed by cutting the profile of a grinding wheel from the highest peak to the
deepest valley at defined intervals from a horizontal plane. Important components of the Abbott-Curve
are the core height Sk, which corresponds to the distance between the highest and lowest levels of the
core of the surface, the reduced peak height Spk, which indicates the mean height of the protruding
peaks above the core height, and the reduced valley height Svk, which indicates the prominent valleys
below the core.
The Abbott-Curve also opens the possibility to calculate the area-related empty volume Vv(p)
of the cavities for a given material proportion p. Based on this, the empty volume of the valleys of
a surface Vvv and the empty volume of the core Vvc is calculable. Analogously to the calculation
of the empty volume, the peak material volume Vmp is calculable for a given material proportion.
However, the Abbott-Curve does not provide information about the ratio of the proportions of bond
and abrasive grains on the surface of a grinding wheel [11].
Weiß, Duscha, and Rasim pursued in their research a different approach for the description of the
grinding wheel topography. They quantitatively described the contact conditions in grinding by means of
a grain engagement model, which bases on the limiting cutting edge offset angle εlim according to Kassen
and Werner [12,13]. Therefore, they developed the software Topo-Tool, which focuses on the analysis and

26
2
Inventions 2017, 2, 34

quantitative description of the topography by analyzing the kinematic cutting edges. Weiß [14] determined
the number of theoretical kinematic grain surfaces Nkin , as well as their respective size, Akin , and calculated
the fractions of these surfaces in the tangential and normal direction to the grinding direction Akin,t and
Akin,n . He quantified the engagement surfaces of measured topographies of ceramic and synthetic resin
bonded grinding wheels depending on their specification, as well as the average tangential and normal
cutting engagement surfaces Ākin,t and Ākin,n as a function of the mean grain diameter [15]. Duscha used
the kinematic engagement surfaces of a grinding wheel to determine the maximum surface pressure in
the grinding process. In this way, he was able to draw conclusions about the real mechanical loads in
the grinding process [16]. Rasim enhanced the Topo-Tool model with the four additional characteristic
values peak angle β, rake angle γ, and sharpening angle δ, as well as the opening angle α to describe the
kinematic cutting engagement surfaces. He determined the frequency distribution of the grain shapes,
which he used to analyze the influence of the shape of the cutting edge surfaces on the energy conversion
in the grinding process. In his investigations, he varied the grain type in galvanic bonded grinding wheels,
and was thus able to model the influence of the grinding wheel topography on the energy conversion in
the grinding process for the first time [17]. However, he did not undertake a comprehensive variation of
the volumetric grinding wheel composition. Figure 1 gives an overview of the potential parameters for
describing the properties of functional surfaces and grinding wheel topographies.
Despite the previous research, the state of the art shows a substantial knowledge deficit about the
influence of the volumetric grinding wheel composition, and the resulting grinding wheel topography
on the thermo-mechanical stress collective acting on the workpiece external zone. Thus, it is not
possible to make a quantitative statement about the influence of the topography geometry on
the thermal and mechanical process load acting on the workpiece external zone of a component.
Consequently, the knowledge about the influence of the topography of a grinding wheel on the
component’s surface and external zone properties is insufficient.
Therefore, the aim of current research is an empirical-analytical model for the prediction of
the thermo-mechanical stress collective as a function of the grinding wheel topography. For this
purpose, a methodology has been developed that enables the prediction of the topography-dependent
thermo-mechanical load in the grinding process.

27
3
Inventions 2017, 2, 34

Figure 1. Parameters for the description of the properties of functional surfaces and grinding
wheel topographies.

2. Materials and Methods


First, grinding tests with multilayered, synthetic resin-bonded cubic boron nitride (CBN) grinding
wheels have been carried out. In the grinding experiments, a tool grinding machine S22P turbo of
the company ISOG was used to investigate the thermo-mechanical stress collective depending on
the grinding wheel topography. During the experimental investigations, 13 synthetic resin-bonded
CBN grinding wheels from the company TYROLIT Schleifmittelwerke Swarovski K.G. (Schwaz,
Austria) were used to grind the bearing steel 100Cr6. The grinding wheels differed in their volumetric
composition, in the average grain size, and in the grain type. The grain types ABN200 and ABN800
from the company Element Six (Burghaun, Germany) were used. The grain type ABN200 is described
as “sharp” with high strength and thermal stability. The grains predominantly have an octahedral

28
4
Inventions 2017, 2, 34

shape and are rounder than the grains of the grain type ABN800. An overview of the used grinding
wheel specifications is shown in Figure 2.

*ULQGLQJ :KHHO6SHFLILFDWLRQV

*UDLQ W\SH*7 $%1 $%1




3RUHSURSRUWLRQ

93



 0HDQ JUDLQ
GLDPHWHU G* —P

   
*UDLQ SURSRUWLRQ 9* 

Figure 2. Overview of the used grinding wheel specifications.

Approximately 160 grinding tests were carried out, each with different experimental conditions to
investigate the influence of the grinding wheel topography on the thermo-mechanical stress collective.
A full-factor experimental design was chosen in order to fully explore the influence of the individual
input factors. By varying only one input variable at otherwise constant input variables, it was possible
to identify the influence of each input variable on the process result. Thus, almost-repeated tests were
carried out in the experimental design, which could be confirmed on the basis of the results. Both, the
composition of the grinding wheels and the dressing process were varied to adjust the topography of
the grinding wheels. The experiments were carried out in an up-grinding process. In the experiments,
the grinding wheel topography was varied by changing the volumetric composition of the grinding
wheels. The CBN abrasive wheels were profiled by a diamond form roller with a mean grain size of
301 μm and were sharpened by means of sharpening blocks of the type A240 H5 V from TYROLIT
Schleifmittelwerke Swarovski K.G. company. They were dressed with different dressing overlap
ratios Ud = 1; 2; 4, a constant dressing feed aed = 2.5 μm and a speed ratio of qd = 0.8. In addition,
the topography was varied by using the sharpening parameters specific sharpening material removal
rate Q’Sb = 100 mm3 /mms and specific sharpening material removal V’Sb = 50, 100 and 150 mm3 /mm.
The grinding parameters grinding wheel circumferential speed vs = 30 m/s and infeed ae = 100 μm were
kept constant. The workpiece speed was varied in three steps between 1.5 m/min < vw < 4.5 m/min
to set the specific material removal rate varied between 2.5 mm3 /mms < Q’w < 7.5 mm3 /mms.
The interpretation of the thermo-mechanical stress during grinding is only possible with a
sufficiently precise knowledge of the occurring grinding force and temperature, as well as the geometric
contact conditions along the real contact arc. Therefore, a temperature-force measuring platform
has been manufactured to measure the occurring grinding forces and temperatures. In order to
determine the thermal loads during the grinding process parallel to the mechanical loads, a temperature
measurement method according to Duscha, based on the work of Choi and Batako, was used [16,18,19].
A constantan foil (55% Cu, 44% Ni and 1% Mn) was connected to a constantan wire with the help of a
TL-WELD-220V thermocouple welding machine from OMEGA. The resulting metallic conductor was
fixed between two screwed workpiece parts. Ceramic mica discs isolated the constantan foil from the
two workpiece parts beside. The downstream workpiece part was connected with an iron conductor

29
5
Inventions 2017, 2, 34

to a reference point. As a result of the passage of the grinding wheel during the grinding process,
the constantan foil was machined, and, due to plastic deformation, a contact with the downstream
part of the workpiece was closed over the isolation. The closed contact between the metal pair enabled
the temperature measurement due to the occurring thermal stress by means of the electrical charge
transfers in a high repeatability and certainty of results.
To ensure that the initial wear of the grinding wheels was overcome before the grinding
experiments, each grinding wheel was ground in after the dressing process. Once the grinding
wheel produced a constant grinding force, the grinding in process was stopped.
In order to identify the influence of the grinding wheel topography on the thermo-mechanical stress
collective, the topography of the grinding wheels was recorded and analyzed by a three-dimensional
(3D) Laser Scanning Confocal Microscope VK-X 150 before each experiment. The analysis of the
topographies was carried out with the analysis module of the company Keyence (Osaka, Japan) and with
the MATLAB-based Topo-Tool developed at the Laboratory for Machine Tools and Engineering (WZL)
of RWTH Aachen University [14]. The analysis module of the laser scanning microscope enables the
determination of the volume, height, distance and function parameters listed in Figure 1 according to
DIN EN ISO 25178 [10]. The Topo-Tool enabled the identification of the kinematic engagement surfaces
of a grinding topography, as well as the calculation of the size of the engagement surfaces, depending
on the process parameters. The calculation is based on the Formula 1 calculated by Kassen and Werner
for the calculation of the limiting cutting edge angle εlim [12]. The calculated limiting cutting edge angle
εlim results from the infeed of the grinding wheel during simultaneous relative movement between the
grinding wheel and the workpiece in the circumferential direction of the grinding wheel. In Figure 3,
the approach for calculating the kinematic engagement surfaces is shown schematically, using the example
of a small grinding wheel cutout.
 1
vw ae 2
ε lim = ·2· (1)
vs ds

Figure 3. Approach for calculating the kinematic engagement surfaces on the topography.

In addition to the identification of the kinematic engagement surfaces, the size of the individual
kinematic engagement surfaces Akin and the proportions in normal and tangential direction Akin,n and
Akin,t were determined, according to [14,16]. Based on this, the mean kinematic engagement surfaces
in normal and tangential direction Ākin,n and Ākin,t were calculated. Furthermore, the analysis of the
shape of the kinematic cutting edges was carried out using the characteristic values of the peak angle
β, the rake angle γ, and the sharpening angle δ [17].

3. Results
The aim of the research was to investigate the influence of the grinding wheel topography on the
thermo-mechanical stress collective. Therefore, first the influence of the listed topography parameters
on the grinding force and the grinding temperature was analyzed. Parameters with an influence on the
thermo-mechanical stress collective were subsequently tested for correlations with each other and were
then grouped. Subsequently, the influence of a change of the relevant topography parameters on the
process results was examined in detail. The high number of experiments and investigated parameters

30
6
Inventions 2017, 2, 34

did not permit any direct analysis and interpretation of all the results. Therefore, first investigations
that are presented in this paper were carried out with a constant specific material removal rate of
Q’w = 2.5 mm3 /mms, a constant dressing overlap ratio Ud = 1 and with grinding wheels with a grain
proportion of VK = 25% and mean grain size of dG = 126 μm.
The analysis of the test results was carried out with the aid of the open-source software
ORANGE, which is a component-based visual programming software package for data analysis,
data visualization, machine learning, and data mining [20]. The visualization of the correlations
between topography parameters and process parameters was carried out using scatter plot matrices in
MATLAB. This enabled the rapid identification of correlations between the parameters.

3.1. Correlations between the Process Result Parameters


Figure 4 shows the correlation between the maximum temperature Tmax and the grinding force in
tangential and normal direction Ft and Fn for the specific material removal rate Q’w = 2.5 mm3 /mms.
In the present case, the highest value for the degree of determination R2 was found with a linear
approach to the relationship between the grinding force and the temperature. In particular, there was
a very strong linear correlation between the maximum temperature Tmax and the grinding forces in
normal and tangential direction Fn and Ft . Based on this finding, the influence of the topography on
the process behavior was investigated by means of the grinding temperature.

Figure 4. Correlation between the process force and the process temperature.

3.2. Correlations between the Topography Parameters


In the following, the characteristic topography parameters, which had a similar influence on the
thermo-mechanical stress collective were investigated and grouped. Three parameter groups were
identified by the analysis, so that a reduction of the total of 33 investigated topography parameters to
three parameter groups could be realized for describing the influence of the grinding wheel topography
on the thermo-mechanical stress collective.
The results did show no correlation between the height parameters Sa, Sz, Sq, Ssk, Sku, Sp,
and Sv among each other and the process results maximum temperature Tmax and the process force
components Fn and Ft . The process results also showed no significant dependencies on the height

31
7
Inventions 2017, 2, 34

parameters Str, Sal, and Std. Therefore, height parameters are not suitable for describing the influence
of the grinding wheel topography on the thermo-mechanical stress collective.
The volume parameter texture aspect ratio of the surface Str, autocorrelation length of the texture
Sal, and the texture direction Std, describe the spatial structure patterns of the topography. Significant
dependencies of the process variables were not found in the investigations, so that the volume
parameters do not seem to be suitable for the quantitative description of the influence of the topography
on the grinding process behavior.
In mechanical engineering, hybrid parameters are used in various applications to describe wear
and tribological behavior. They combine spatial and vertical parameters. The developed interfacial
area ratio Sdr indicates the ratio between the true surface of the topography and the projected base
area in the X–Z plane. Thus, it is a measure of the three-dimensional roughness of the analyzed
surface. A perfectly planar surface has a value of Sdq = 0. The arithmetic mean value of the peak
curvature Spc, the developed interfacial area ratio Sdr, and the root mean square gradient Sdq showed
a particularly strong correlation. The correlation between the arithmetical mean value of the peak
curvature Spc and the developed interface area ratio Sdr was best describable by a quadratic function
with a coefficient of determination of R2 = 0.9589, which corresponds to an excellent correlation
(see Figure 5). The correlations between the above-mentioned parameters and the process variables
were investigated and they showed identical trends. For the further procedure, the three parameters
were grouped and the influence of the arithmetical mean value of the peak curvature Spc was analyzed
as representative for the parameter cluster.

Figure 5. Correlation between the developed interfacial area ratio, the arithmetic mean peak curvature
and the root mean square gradient.

Similarly, there were strong correlations between the functional parameters describing the upper
part of the Abbott curve, see Figure 6. Since the correlation coefficients between the reduced peak
height Spk, the peak extrem heigt Sxp, and the peak material volume Vmp assume approximately a
coefficient of determination of nearly R2 = 1, an almost perfect linear relationship exists between the
topography characteristics. Here, the formation of a cluster is obvious, since all the parameters also
had the same relation to the process variables.
The angles of the kinematic cutting edges, which were identified by the Topo-Tool, also correlated
strongly. Figure 7 shows approximately linear relationships between the angles of the kinematic
cutting edge areas. When the mean peak angle β or the mean sharpening angle δ increase, the mean

32
8
Inventions 2017, 2, 34

rake angle γ decreases. Here, too, a grouping of these parameters was appropriate, in order to reduce
complexity for the analysis of the influence of the grinding wheel topography on the thermo-mechanical
stress collective.

Figure 6. Correlation between the peak material volume, the reduced peak height and the peak
extreme height.

Figure 7. Correlations between the angles of the kinematic cutting edge areas.

3.3. Influence of the Topography Parameters on the Chip Thickness and Process Result Parameters
Based on the rapid identification of correlations between the topography parameters and the
process parameters, a detailed analysis of the influence of the relevant topography parameters on
the thermo-mechanical stress collective was carried out. Figure 8 shows the correlations between the
hybrid parameters and the process parameter mean maximum chip thickness hcu, max . The maximum
chip thickness was investigated by means of the Topo-Tool, which measures the biggest vertical
distance between the lowest and the highest points of the tangential kinematic engagement area, as
described in Figure. Between the developed interface ratio Sdr and the process parameters Fn , Ft ,
and Tmax , only weak correlations were identified. An influence of these surface parameters on the
process behavior is physically close, but due to the coefficient of determination of only R2 = 0.62, a low

33
9
Inventions 2017, 2, 34

suitability of these topography characteristics is assumed to describe the influence of the topography
on the thermo-mechanical stress collective solely.
The further investigated functional parameters are derived from the Abbott curve,
which characterizes the vertical material distribution of a surface structure. The Abbott curve is created
by applying the material portion above the profile height. Functional parameters, such as the reduced
dale height Svk, or the core material volume Vmc, which describe the geometry and structure of the
grooves on a surface, did not influence the force or temperature in the grinding process. In particular,
the reduced dale height Svk, which evaluates surfaces with regard to their ability to absorb and to
serve fluid to adjust the lubricating properties in tribosystems, did not show any correlation with force
and temperature, contrary to expectations.

Figure 8. Correlation between the hybrid parameters and the mean max. chip thickness.

However, topography parameters that describe the characteristics of the peaks of a topography
showed correlations with the thermo-mechanical stress collective. The peak extreme height Sxp is the
difference of heights at the areal material ratio values p = 0% and q = 50%. The corresponding material
portions p and q correlated strongly with the grain protrusion. Influences on the maximum temperature
Tmax and the occurring grinding forces Ft and Fn were measured. As the grain protrusion increased
and the value of Sxp raised, the grinding force and the temperature dropped. The same applied to the
relationship between the reduced peak height Spk, which corresponds to the reduced height of the
peaks above the core surface, and the grinding forces in the normal and tangential directions.
The peak material volume Vmp describes the material volume for an areal material ratio value
of 10%. It corresponds to the volume, which is occupied by the tips of the topography. The larger
the volume of the tips, the higher is the protrusion of the grains, or the amount of grains out of the
core volume of the topography. If the peak material volume Vmp is small, then the core volume
Vmc is correspondingly larger. This is the case if grinding wheels are smooth with a small grain
protrusion. In the case of a small grain protrusion, only a small quantity of grinding fluid can enter the
contact zone, which influences the contact conditions. As the peak material volume Vmp increased,
the tangential grinding force Ft , as well as the maximum temperature Tmax , dropped significantly,
as presented in Figure 9. The coefficient of determination is in the range of R2 = 0.5–0.6. The relatively
small R2 indicates that the parameters Sxp, Vmp, and Spk influence the thermo-mechanical stress
collective, but are likewise not suitable for describing the influence of the grinding wheel topography
on the thermo-mechanical stress collective solely.

34
10
Inventions 2017, 2, 34

Figure 9. Correlation between volume and functional topography parameters and the process results.

The feature parameters include the peak density of the surface Spd and the arithmetic mean peak
curvature Spc. The peak density Spd indicates the number of roughness peaks of a topography relative to
the considered area [10]. This parameter ignores that a large part of the identified tips does not come into
engagement because of shading effects during the grinding process. Accordingly, the meaningfulness
of this parameter is very low. With regard to the process results, Spc showed a correlation with the
maximum grinding force in normal and tangential direction Fn and Ft , as well as to the maximum
temperature Tmax . The results showed distinctive trends, but the coefficients of determination R2 were
low. So the feature parameters were not taken into account for detailed investigations. The test results
showed a linear correlation between the mean tangential kinematic engagement area Ākin,t , and the mean
maximum chip thickness hcu, max . This correlation becomes already clear from geometrical considerations.
A similar, but lower effect showed an increase of the total tangential kinematic engagement area Akin,t .
The relationship between the average tangential kinematic engagement area Ākin,t and the averaged
maximum chip thickness hcu, max is described by a linear mathematical function, with a coefficient of
determination of R2 = 0.6463 (see Figure 10).
The mean peak angle β, the mean rake angle γ and the mean sharpening angle δ showed a
strong influence on the averaged maximum chip thickness hcu, max . The coefficient of determination
for a linear relationship between β and hcu, max was about R2 = 0.69, which corresponds to a good
regression (see Figure 11). Direct connections between the averaged maximum chip thickness hcu, max
and the grinding force, as well as the temperature were not detectable. One possible reason is that the
calculation of hcu, max solely does not take the different grain protrusions during the investigations
into account.

35
11
Inventions 2017, 2, 34

Figure 10. Correlation between kinematic engagement grain area and chip thickness.

Figure 11. Correlation between kinematic engagement grain angles and the mean maximum chip thickness.

In Figure 12 the relationship between the mean rake angle γ, the developed interfacial area ratio Sdr,
and the mean tangential kinematic engagement area Ākin,t is shown. A strong correlation prevails between
the mean sharpening angle δ and the mean tangential kinematic engagement area Ākin,t . An increase in
the developed interfacial area ratio Sdr is associated with a rising mean tangential kinematic engagement
area Ākin,t . Based on these correlations, this means that the mean grain surface contact angle β, γ, and δ
as a group, and the topography characteristics Spc, Sdr, and Sdq as a group correlate as well.

36
12
Inventions 2017, 2, 34

Figure 12. Correlation between the mean tangential kinematic engagement area, the mean sharpening
angle and the developed interfacial area ratio.

The results of the empirical investigations showed that a change of each identified topography
parameter group had an influence on the formation of the thermo-mechanical stress collective.
However, the results of the investigation also allowed for the conclusion that the thermo-mechanical
load in the grinding process is caused by a superposition of the influences of these identified topography
parameters. Therefore, the identified topography parameters were investigated in combination.
In this context, Figure 13 shows the measured maximum grinding temperature Tmax depending
on the peak material volume Vmp and the mean peak angle β. The results showed that the maximum
grinding temperature Tmax decreased strongly with an increasing mean peak material volume Vmp.
However, the results show that topographies with a similar peak material volume Vmp did not necessarily
lead to a similar maximum grinding temperature Tmax . By a combined examination of the mean peak
material volume Vmp and the mean peak angle β in Figure 13, it can be seen that the fluctuating maximum
grinding temperatures at almost constant Vmp can be attributed to the geometry of the kinematic peak
angles. If, for example, the mean peak angle β increased, the maximum grinding temperature Tmax
increased significantly with a constant peak material volume Vmp. This was especially noticeable at
low values of the mean peak material volume Vmp. An increase of the mean peak angle from β = 75◦
to β = 95◦ at a constant mean peak material volume of Vmp = 0.4 led to an rising maximum grinding
temperature Tmax from nearly 500 ◦ C to more than 650 ◦ C. With increasing Vmp, the influence of the angle
decreased. It can be assumed that the influence of the mean peak angle β decreased with increasing Vmp,
since additional chip space allowed for an increased dissipation of the process heat. The dependence of
the grinding temperature on the grinding wheel topography could be approximated with the help of
the two parameters with a coefficient of determination of R2 = 0.71 and a root mean square error of 85.2,
which corresponds to a good accuracy.

37
13
Inventions 2017, 2, 34

Figure 13. Grinding temperature depending on the topography parameters peak material volume and
mean peak angle.

4. Discussion of the Results


Based on the presented evaluations, it can be summarized that three groups, which correlated
with the temperature and force during the grinding process, with each three characteristic parameters
were identified. The first group, consisting of the developed interface area ratio Sdr, the arithmetic
mean peak curvature Spc, and the root mean square gradient Sdq describes the shape of the roughness
on the grinding wheel surface. An increase of these parameters resulted in a reduction of the normal
grinding force Fn and the maximum grinding temperature Tmax . The second topography parameter
group, consisting of the grain engagement angles mean peak angle β, the mean sharpening angle
δ, and the mean rake angle γ, as well as the kinematic grain surface area in tangential and normal
direction Akin,t and Akin,n showed a strong correlation with the mean maximum chip thickness hcu, max .
In addition, they slightly correlated with the first named group. The third group of topography
parameters includes functional parameters, which describe the material proportions of the topography
by means of the abbott-curve; the reduced peak height Spk, the peak material volume Vmp, and the
peak extreme height Sxp. These topography parameters correlated strongly with each other, but did
not show any correlation with other parameters. A change of these parameters had a very strong
influence on the grinding forces and the temperature. Figure 14 shows a summary of the previously
identified groups, which have a strong influence the thermo-mechanical stress collective.

Figure 14. Overview about the topography parameter groups with an influence on the process behavior.

By a combined consideration of the influence of these three groups, the thermal loads in the grinding
process were approximated with a high degree of accuracy. Thus, the maximum measured grinding
temperature Tmax could be approximated with the help of the parameters peak material volume Vmp
and the mean peak angle β with a coefficient of determination of R2 = 0.71, which corresponds to a good

38
14
Inventions 2017, 2, 34

value, due to the high complexity in grinding technique. The research results show that by means of the
identified topography parameters, it was possible to describe the influence of the topography’s geometry
on the thermo-mechanical stress collective based on only few geometrical parameters quantitatively.
Further investigations should confirm and extend these results.

5. Conclusions
The aim of the investigations was to identify the influence of the grinding wheel topography on
the thermo-mechanical stress collective during the grinding process, and to determine the topography
parameters by which the influence of the topography on the thermo-mechanical stress collective can be
described quantitatively. For this purpose, empirical-analytical examinations with different grinding
wheels were performed. A high variation of topographies was realized by means of 13 different
grinding wheel specifications and by varied dressing and sharpening strategies.

• During empirical investigations, topography parameters could be identified and classified into
groups that had an influence on the thermo-mechanical stress collective
• Each group combines topography parameters, which had the same influence on the formation of
the thermo-mechanical stress collective (Figure 14).
• The isolated consideration of these groups, however, did not provide a clear explanation of
the dependence of the thermal and mechanical loads in the grinding process on the grinding
wheel topography.
• The combined consideration of the groups around the topography parameters peak material
volume Vmp and mean peak angle β allowed for the quantitative description of the dependence
of the thermo-mechanical stress collective on the geometry of the grinding wheel topography.
So, it became possible to describe the influence of the topography on the stress collective during
grinding by a few topography parameters for the first time quantitatively.

In further investigations, the influence of the dressing overlap ratio, as well as the volumetric
composition and the related machining volume on the grinding wheel topography and its kinematic
components, as well as on the grinding process result will be investigated.

Acknowledgments: The author would like to thank the German Research Foundation (Deutsche
Forschungsgemeinschaft—DFG) for the support of the depicted research within the project KL500/178-1
“Influence of the grinding wheel topography on the thermo-mechanical stress collective”. Special thanks to
the company TYROLIT Schleifmittelwerke Swarovski K.G., for the intensive discussion and for their support in
providing specimens and grinding wheels in a variety of specifications.
Author Contributions: Sebastian Barth conceived, designed and performed the experiments, analyzed the data,
interpreted the results and wrote the manuscript. Fritz Klocke contributed to the discussion and reviewed and
approved the manuscript.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Schmitt, R. Abrichten von Schleifscheiben mit Diamantbestückten Rollen. Ph.D. Thesis, Technische Universität
Carolo-Wilhelmina zu Braunschweig, Braunschweig, Germany, 1968.
2. Darafon, A.; Warkentin, A.; Bauer, R. Characterization of grinding wheel topography using a white chromatic
sensor. Int. J. Mach. Tools Manuf. 2013, 70, 22–31. [CrossRef]
3. Frühling, R. Topographische Gestalt des Schleifscheibenschneidenraumes und der Werkstückrautiefe beim
Außenrundeinstechschleifen. Ph.D. Thesis, Technische Universität Carolo-Wilhelmina zu Braunschweig,
Braunschweig, Germany, 1976.
4. Paucksch, E.; Holsten, S.; Linß, M.; Tikal, F. Zerspantechnik: Prozesse, Werkzeuge, Technologien, 12th ed.;
Studium, Vieweg+Teubner Verlag/GWV Fachverlage GmbH: Wiesbaden, Germany, 2008.
5. Uhlmann, E.; Koprowski, S.; Weingaertner, W.L.; Rolon, D.A. Modelling and Simulation of Grinding
Processes with Mounted Points. Procedia CIRP 2016, 46, 599–602. [CrossRef]

39
15
Inventions 2017, 2, 34

6. Nowicki, B. Multiparameter representation of surface roughness. Wear 1985, 102, 161–176. [CrossRef]
7. Hecker, R.L.; Ramoneda, I.M.; Liang, S.Y. Analysis of wheel topography and grit force for grinding process
modeling. J. Manuf. Process. 2003, 5, 13–23. [CrossRef]
8. Nguyen, A.T.; Butler, D.L. Correlation of grinding wheel topography and grinding performance. J. Mater.
Process. Technol. 2008, 208, 14–23. [CrossRef]
9. Hübert, C.; Hahmann, D.; van der Meer, M.; Hahmann, W.C.; Pekárek, M.; Rickens, K.; Mauren, F.;
Mutlugünes, Y. Charakterisierung von Schleifscheibentopographien aus fertigungstechnischer Sicht.
Diamant Hochschulwerkzeuge 2009, 4, 40–47.
10. Deutsches Institut für Normung e.V. (DIN). Geometrische Produktspezifikation (GPS)—Oberflächenbeschaffenheit;
17.040.30 (DIN EN ISO 25178); Beuth Verlag GmbH: Berlin, Germany, 2013.
11. Duscha, M.; Klocke, F.; Wegner, H.; Gröning, H. Erfassung und Charakterisierung der
Schleifscheibentopographie für die anwendungsgerechte Prozessauslegung. Diam. Bus. 2009, 28, 28–33.
12. Kassen, G. Beschreibung der Elementaren Kinematik des Schleifvorganges. Ph.D. Thesis, RWTH Aachen
University, Aachen, Germany, 1969.
13. Werner, G. Kinematik und Mechanik des Schleifprozesses. Ph.D. Thesis, RWTH Aachen University, Aachen,
Germany, 1971.
14. Weiß, M.; Klocke, F.; Barth, S.; Rasim, M.; Mattfeld, P. Detailed Analysis and Description of Grinding Wheel
Topographies. J. Manuf. Sci. Eng. 2017, 139, 54502. [CrossRef]
15. Weiß, M. Einfluss der Spezifikation Mehrschichtiger CBN-Schleifwerkzeuge auf das Schleifprozessverhalten:
Specification Influence of Multilayer CBN-Grinding Wheels on the Grinding Process Behaviour, 1st ed.;
Apprimus Verlag: Aachen, Germany, 2016.
16. Duscha, M. Beschreibung des Eigenspannungszustandes beim Pendel- und Schnellhubschleifen. Ph.D. Thesis,
RWTH Aachen University, Aachen, Germany, 2014.
17. Rasim, M. Modellierung der Wärmeentstehung im Schleifprozess in Abhängigkeit von der Schleifscheibentopographie,
1st ed.; Apprimus Verlag: Aachen, Germany, 2016.
18. Choi, H.-Z. Beitrag zur Ursachenanalyse der Randzonenbeeinflussung beim Schleifen. Ph.D. Thesis,
Technische Universität Hannover, Hannover, Germany, 1986.
19. Batako, A.D.; Rowe, W.B.; Morgan, M.N. Temperature measurement in high efficiency deep grinding. Int. J.
Mach. Tools Manuf. 2005, 45, 1231–1245. [CrossRef]
20. Demsar, J.; Curk, T.; Erjavec, A.; Gorup, C.; Hocevar, T.; Milutinovic, M.; Mozina, M.; Polajnar, M.; Toplak, M.;
Staric, A.; et al. Orange: Data Mining Toolbox in Python. J. Mach. Learn. Res. 2013, 14, 2349–2353.

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

40
16
inventions
Article
Influence of Cutting Speed on Subsurface Damage
Morphology and Distribution in Ground Fused Silica
Georg Schnurbusch 1, *, Ekkard Brinksmeier 2 and Oltmann Riemer 1
1 Laboratory for Precision Machining (LFM), University of Bremen, Badgasteiner Str. 2, 28359 Bremen,
Germany; [email protected]
2 Foundation Institute of Materials Science, Badgasteiner Str. 3, 28359 Bremen, Germany;
[email protected]
* Correspondence: [email protected]; Tel.: +49-(0)421-218-51141

Received: 3 July 2017; Accepted: 4 August 2017; Published: 10 August 2017

Abstract: In optical fabrication, brittle-hard materials are used for numerous applications. Especially
for high-performance optics for laser or lithography applications, a complex and consistent
production chain is necessary to account for the material properties. Particularly in pre-processing,
e.g., for shaping optical components, brittle material behavior is dominant which leads to a rough
surface layer with cracks that reach far below the surface. This so called subsurface damage (SSD)
needs to be removed in subsequent processes like polishing. Therefore, it is essential to know the
extent of the SSD induced by shaping for an efficient design of precise corrective processes and for
process improvement. Within this work the influence of cutting speed on SSD, in fused silica, induced
by grinding has been investigated. To analyze the subsurface crack distribution and the maximum
crack depth magnetorheological finishing has been appointed to polish a wedge into the ground
surface. The depth profile of SSD was analyzed by image processing. For this purpose a coherent
area of the polished wedge has been recorded by stitching microscopy. Taking the form deviation
of the ground surface in to account to determine the actual depth beneath surface, the accuracy of
the SSD-evaluation could be improved significantly. The experiments reveal a clear influence of the
cutting speed on SSD, higher cutting speeds generate less SSD. Besides the influence on the maximum
crack depth an influence on the crack length itself could be verified. Based on image analysis it was
possible, to predict the maximum depth of cracks by means of crack length.

Keywords: optics manufacturing; precision grinding; subsurface damages; brittle hard materials

1. Introduction
Grinding is still an indispensable process step in the manufacturing of high precision optics, due
to its high removal rates and well-controlled shaping. However, due to the prevailing material removal
mechanisms, the grinding of brittle materials inevitably leads to microscopic cracks beneath the surface.
This so called subsurface damage (SSD) influences the operability and lifetime of high-performance
optics like semiconductor or high power laser applications [1,2]. The shaping process by grinding is
followed by various corrective processes like polishing for smoothing the surface, improving the shape
accuracy and removing SSD. The maximum crack depth at each process step decisively determines
the extent of the polishing and corrective processes required and thus the costs of the overall optic
manufacturing process [3]. According to this, a main issue in the production of high performance
optics is to minimize the process-related SSD in grinding.
The undefined cutting edge and brittle material removal behavior make grinding a stochastic
process. Single grains with high protrusions lead to a local overload of the normal force and cause
the deepest cracks [4]. For fine-grained diamond grinding wheels the influencing variables grain
protrusion and number of active grains cannot be determined directly, therefore it is nearly impossible

Inventions 2017, 2, 15; doi:10.3390/inventions2030015 41 www.mdpi.com/journal/inventions


Inventions 2017, 2, 15

to model the grinding process or to predict SSD. Thus, the process design for optics manufacturing still
depends on empirical data. To date, the grain size serves as an orientation [5]. The stability of grinding
processes as well as the accuracy in ascertaining the maximum depth of SSD is essential for increasing
the efficiency in optics manufacturing. The same counts for the investigation of the influencing
variables on the process-induced SSD. The accuracy of the evaluation is even more important for fine
grinding where crack depths are only a few micrometers.
Nature and extent of subsurface damage in glasses makes quantitative assessment of damage
difficult [6]. To date, it is still challenging to determine subsurface damage in optical glasses.
Since the measurement of SSD with destructive methods is complex as well as cost-intensive
and time-consuming, several attempts were undertaken to correlate process parameters and
measurable output values, like surface roughness. The results of Lambropoulos et al. and Wang
et al. show, however, a broad distribution in the results of these methods [7,8]. Even though there are
numerous non-destructive approaches to determine the crack depths beneath the machined surface,
the destructive methods still provide the most accurate results [7]. Over the past few years, the
combination of magnetorheological finishing (MRF) and etching has proven to be the most appropriate
strategy [7,8]. MRF provides a highly deterministic material removal process without generating
new or propagating existing SSD [9]. The subsequent etching process serves to open the cracks on
the polished surface to provide a better contrast in optical imaging. The damage free polishing of a
wedge allows for a glance under the ground surface. Thus, not only the maximum crack depth can be
determined, but also the density and shape of cracks can be observed at different depths. Resolution
of this analysis obviously depends on the wedge gradient, as the area increases with decreasing
wedge angles.
Simple investigations of SSD by MRF technique are limited to the determination of the maximum
crack depth [10,11]. Miller et al. first analyzed the depth profile of the crack density by single
micrographs along the wedge [4]. To enable analysis for a larger wedge area, adjacent microscope
images were stitched by using a computer-based microscope system with a driven stage [8]. This method
offers the great advantage, to analyze the recorded data by image processing. The SSD depth is
calculated from the difference in height between the profiles of the ground surface and the wedge
surface. In previous publications, a perfect initial surface was assumed, however, even fine grinding
generates form deviations of a few micrometers. For rough cuts and pre-shaping this seems negligible,
but in the case of fine grinding these deviations can be on the same scale as the SSD.
Within this work, the influence of cutting speeds on SSD for ground fused silica was investigated.
To analyze the maximum crack depth and crack distribution MRF technique was used in conjunction
with dry etching. Finally, for the first time, the initial surface profile was considered to enhance the
accuracy to determine the actual crack depth.

2. Materials and Methods


Fused silica samples with a diameter of 100 mm were ground on an ultraprecision 5-axis grinding
machine (Nanotech 500 Freeform Generator, Moore Nanotechnology Systems). The experiments
comprise a variation of cutting speeds between 30 and 62 m/s, in five subdivisions, with three
repetitions (Table 1). For process monitoring, the grinding forces were measured by a multicomponent
dynamometer (CompactDyn, Kistler, Winterthur, Switzerland) mounted beneath the hydrostatic
grinding spindle. The grinding kinematics is shown in Figure 1.
A resin bonded diamond grinding wheel with a diameter of 100 mm, D20 (C70) was applied.
All samples were ground with a constant rotary worktable speed nW = 50 rpm and radial feed
vfr = 10 mm/s. Based on a fine grinding process, the material removal was conducted in four
passes with 20 μm and one final pass with 10 μm depth of cut. In order to guarantee a uniform
initial state, all samples were pre-ground to plane shape before running the actual experiments
(pre-processing, ae = 100 μm; four passes with 25 μm). Furthermore, the grinding wheel was dressed

42
Inventions 2017, 2, 15

and balanced before each grinding experiment. This ensures that the measured SSD results from the
actual process parameters.

(a) (b)

Figure 1. (a) Grinding kinematics—cross grinding with rotating workpiece; (b) Machine setup.

Table 1. Process parameters.

Cutting Speed Depth of Cut ae Worktable Radial Feed vfr


Process Step
vc (m/s) (μm) Speed nW (rpm) (mm/min)
1st step: pre-processing 30 100 (4 × 25) 50 10
2nd step: SSD experiments 30–62 90 (4 × 20, 1 × 10) 50 10

3. Results

3.1. Grinding Force and Surface Roughness


Based on the constant rotary worktable speed, the feed rate decreases continuously from edge
to center. Since the remaining parameters are kept constant, this leads to a reduction in material
removal rate and thus a reduction in chip thickness. Consequently, normal forces (Figure 2a) and
roughness decrease from edge to center. In conventional grinding, it is generally assumed that higher
cutting speeds lead to a lower chip thickness and thus also to lower grinding forces [12]. In the
experiments presented in this work, however, the opposite is shown. Larger cutting speeds led to
higher normal forces (Figure 2b). For brittle materials like fused silica, the removal mechanisms
depend on the chip thickness. Due to a higher rate of ductile material removal, the forces increase with
the cutting speed [13]. To make the results comparable and regarding the position of the maximum
crack depth cmax at the MRF wedge, the presented values were obtained at a distance of 15 mm from
workpiece center.

(a) (b)

Figure 2. (a) Typical force curve (Fn ) at vc = 62 m/s) for constant worktable speed nW ; (b) Normal
forces for different cutting speeds (distance from workpiece center 15 mm).

43
Inventions 2017, 2, 15

The surface roughness was measured by white light interferometry (WLI, Talysurf CCI HD, Taylor
Hobson) with 20× magnification. Along circles of different radii six spots with an area of 840 × 840 μm2
were measured with an angular distance of 60◦ . To exclude the waviness of the surface a long pass
filter was set to 250 μm.
The results of the roughness measurement (e.g., in Figure 3) show an opposite trend towards
normal force. This agrees well with the assumption that the proportion of ductile material behavior
increases with higher cutting speeds.

(a) (b)

Figure 3. (a) White light interferometer (WLI) measurement (20× magnification): 2D-surface roughness
(vc = 30 m/s, Sa = 160 nm); (b) Surface roughness Sa for different cutting speeds (distance from
workpiece center 15 mm) with linear fit.

3.2. Subsurface Damage Evaluation


For the evaluation of subsurface damage, the micrographs of the MRF wedge need to be merged
with the depth profile. In order to maintain a high accuracy, the samples were provided with a
fiducial mark by laser engraving after grinding. It is thus possible to align the samples for different
measurements and to link the data of tactile measurements and microscope images. Figure 4 shows
the wedge geometry and the entire evaluation procedure.

(a) (b)

Figure 4. (a) Position and geometry of the magnetorheological finishing (MRF) wedge; (b) Sample
processing chain.

After MRF preparation, the wedge surfaces were dry etched. Compared to wet etching the
material removal by dry etching can be controlled with a sub-micron precision [14]. The micrographs

44
Inventions 2017, 2, 15

of the wedge surface were taken with a ZEISS Axio Imager.Z2 Vario with 500× magnification.
750 micrographs were stitched together to an interrelated area of 40 × 1 mm2 . To minimize
contamination from particles the samples were cleaned in an ultrasonic bath and microscopy was
carried out under clean room conditions (ISO 7). In order to determine the actual depth of the wedge
as accurately as possible, the distance between the ground surface (tactile measurement I) and the
wedge surface (tactile measurement II) was calculated from the two tactile measurements.
Due to the form deviation, differences between the polished wedge surface and the actual depth
profile of up to 2 μm (Figure 5) occur in the individual experiments. This is particularly relevant for
fine grinding processes with low SSD. Within the experiments, the differences correspond to 30–40%
of the measured maximum crack depth.

Figure 5. Tactile measurement of ground surface, MRF wedge and depth profile.

The number of cracks decreases with depth along the MRF wedge. With increasing depth of the
wedge, more and more cracks are removed by polishing until the last fracture event can be determined
as the maximum crack depth in the micrographs (Figure 6).

Figure 6. Sketch of subsurface damage (SSD) appearance along the MRF wedge.

To determine the maximum crack depth cmax the lateral position of the last defect on the
micrograph of the polished and etched wedge surface is merged together with the corresponding
information of the calculated depth profile (Figure 7).

45
Inventions 2017, 2, 15

Figure 7. Maximum crack depth depending on the cutting speed.

3.3. SSD Analysis by Image Processing


Besides the evaluation of the maximum crack depth, the micrographs of the MRF wedge provide
the opportunity to examine the crack distribution with increasing depth. Thus, a more extended
examination on the influence of process parameters on grinding results is possible. The high number
of cracks resulting from the size of the inspected wedge area requires a software-supported evaluation.
Therefore the open source software, ImageJ, has been utilized. The software is a highly popular tool
in biology in a wide range of applications and provides a segmentation algorithm. This enables the
analysis of the position and the geometrically relevant data for each individual crack on the stitched
microscope image. After binarization of the images, the cracks appear as uniform black lines on a white
background. To specify the crack length, a best fit ellipse is applied to every single crack (Figure 8).

Figure 8. SSD-Image Processing (ImageJ): (a) raw microscopy data; (b) binary conversion;
(c) determination of crack length L by best fit ellipse.

On average, 150,000 cracks were evaluated for each sample with this method. Figure 9 shows
the ratio of the number of cracks of different lengths to the total crack number (relative frequency) for
different cutting speeds.

46
Inventions 2017, 2, 15

Figure 9. Relative frequencies of the crack length L (filtered data).

After smoothing by a Savitzky–Golay filter, it is easy to see that the distributions differ. As the
cutting speed increases, the amount of shorter cracks (Figure 9a) increases and the amount of longer
cracks decreases (Figure 9b).

4. Discussion

4.1. Maximum Crack Depth


Typically, the crack density decreases along the wedge, i.e., with increasing depth, following a
power law function [15]. At the scale of the surface roughness, the individual cracks intersect and
Suratwala et al. describe this as the rubble zone. A few microns below the surface, the cracks can
be detected individually. Within our experiments, with the given MRF wedge geometry, the lateral
distance between the last cracks could be up to 500 μm. This underlines the importance of size and the
slope of the polished wedge.
After an initial decrease in the maximum crack depth with an increasing cutting speed, a seemingly
constant depth is obtained above 45 m/s (Figure 7). Besides chip thickness reduction, higher spindle
speeds and resulting higher cutting speeds lead to stronger vibrations [16]. This may explain why
there is no further improvement of the maximum SSD-depth at cutting speeds higher than 45 m/s.
The single-grain load, as the main cause for SSD, seems to be confirmed. In theory, the influence of the
material removal rate and thus process parameters such as cutting speed, feed rate and infeed can
be easily assessed. In practice, however, the stochastic nature of the grinding wheel as well as other
influences such as limited machine rigidity have influence on the maximum crack depth.
Miller et al. have shown that the addition of a small amount of larger grains to a fine slurry
for polishing leads to a significant increase in the measured crack depth [4]. The normal force is
distributed to a smaller number of particles than in a homogeneous particle distribution and the
maximum individual grain loads are increased. In the grinding process, the effective grain size
corresponds to the individual grain protrusion, which directly depends on the grain size.
The individually polished and etched wedge surfaces differ considerably, in crack distribution.
Figure 10 shows samples with low and high maximum crack depths. The sample with low SSD shows
a rather arbitrarily arranged crack pattern (Figure 10a), whereas in case of significantly larger SSD
(Figure 10b), the ordered crack formation is particularly remarkable. Large protrusion of individual
grains leads to higher individual loads. As a result, it is not possible to assess the process results by
means of maximum SSD only in conjunction with the cutting speed e.g., vibrations or variation in
grain protrusion.

47
Inventions 2017, 2, 15

(a) (b)

Figure 10. Micrograph of the MRF wedge surface after etching: (a) sample with low SSD (vc = 54 m/s;
cmax = 4.6 μm); (b) sample with larger SSD (vc = 30 m/s; cmax = 6.8 μm).

As stated above, the maximum crack depth cmax results from one particular or only a few grains.
It is thereby subject to a certain variation by the nature of the microstructure of the grinding wheel
and the grinding process. Both, crack depth and crack length, depend directly on the grain load.
The maximum crack depth cmax should therefore also lead to the maximum crack length Lmax . Within
the experiments Lmax shows a linear decay with depth. Figure 11 shows Lmax at intervals of 50 μm
along the polished wedge. To avoid the influence of outliers by intersecting cracks on the linear fit,
the double median crack length LMdn is set as cut-off criteria.

Figure 11. Maximum crack lengths Lmax as a function of depth (vc = 54 m/s) with linear fit.

The zero point of the fit provides a very good correlation with the measured maximum crack
depths (Figure 12).

Figure 12. Measured and calculated maximum crack depths.

48
Inventions 2017, 2, 15

4.2. Fracture Mechanics


Determination of the exact surface topography of a grinding wheel is difficult, even for
coarse-grained tools. Furthermore, the surface changes continuously by wear or is completely reshaped
by dressing. Therefore, attempts are often made to simplify the grinding process by single grain scratch
behavior. The basis for most of the considerations is the work of Lawn, who investigated the cracking
mechanisms of brittle materials under indentation [17]. His experiments showed the formation of
median and lateral cracks as a function of loading and unloading on the workpiece surface. He showed
that the maximum radial crack length cr correlates with the load on the indenter, i.e., P~cr 3/2 [18].
 2/ 3
χr P
cr = with χr ∼ E/H (1)
KIc

With the indentation constant χr (0.046 for fused silica [19]), indenter load P, fracture toughness
KIc , hardness H and Young s modulus E.
Due to the stochastic nature of the microtopography of grinding wheels, loads on a single grain
can not be determined. Equation 1, nevertheless, reveals the maximum grain load to be calculated
from the maximum radial crack length cr , which corresponds to the maximum crack depth cmax .

KIc cmax 3/2


P= (2)
χr

Using equation 2 and the measured maximum crack depths (Figure 7), the maximum individual
grain loads P are about 0.2 N for the grinding wheel applied in this work. This corresponds to
approximately 5% of the measured average normal force. In addition to the evaluation of the maximum
crack depth, the wedge method also makes it possible to determine the crack length L on the polished
surface. Based on the Hertzian indentation mechanics Miller et al. found a correlation between average
crack length and max crack depth [4].
Assuming that the diamond grains are idealized spherical abrasives the circles of contact are
defined by the Hertzian contact zone, with the contact circle radius a.
 1/3
4k
a= Pr (3)
3E

With the elastic mismatch factor k, the applied load P and the radius of the abrasive r. The factor
k relates to the different Young s moduli and Poisson s ratio ν of the workpiece and indenter material,
where the primes refer to the indenter.

9   E 
1 − ν
2
k= 1 − ν2 + (4)
16 E’

In contrast to static indentation, a sliding indenter causes an asymmetry in the tensile stress
field. Since the tensile stresses are relatively small near the front end of the indenter the likelihood of
cone cracks increases in this region. This also predicts that the crack may not completely encircle the
entire contact radius [20]. Miller et al. assumed that the crack propagates over only a quarter of the
contact circumference.  1/3
2πa π 4 k
L≈ = Pr (5)
4 2 3 E’
Finally, the calculated crack length Lcalc as a function of cmax according to Equations (2) and
(5) shows appropriate accordance with the determined average crack length LMdn from the optical
evaluation (Figure 13).

49
Inventions 2017, 2, 15

Figure 13. Measured and calculated crack length for different cutting speeds.

5. Conclusions
To date there is no alternative to the destructive techniques for the evaluation of SSD in optical
glasses. By using modern stitching microscopes and image processing, the MRF wedge method
offers the highest accuracy in the determination of crack distribution and maximum crack depth.
Furthermore, a thorough investigation of the removal mechanisms of the grinding process of optical
glasses is enabled.
Within this work subsurface damage (SSD) was directly determined by the MRF wedge technique
for a series of fused silica ground with different cutting speeds from 30 to 62 m/s. To improve
the accuracy of the SSD evaluation method wet etching was replaced by dry etching. Furthermore,
the form deviation of the ground surfaces was taken into account to determine the actual crack depth.
Thus, a significant influence of the cutting speed on SSD for a fine grinding process, as it is found in
process chains of optical fabrication, could be demonstrated. With stitching microscopy and image
processing the gapless evaluation of the density of cracks and the distributions of crack lengths along
the wedge surface has been implemented. Thus, a dependency of the crack length distribution on the
cutting speed could be confirmed. Based on the model of Miller [4], the relation between calculated
crack length Lcalc as a function of cmax and median crack length LMdn could be demonstrated for SSD of
only a few microns. Furthermore, a relationship between the maximum crack length as a function of
depth and the maximum crack depth could be found. This and the different crack patterns of the SSD
samples supports the thesis that the maximum crack depth is caused by a few individual grains that
also produce the largest crack length.

Author Contributions: Georg Schnurbusch, Ekkard Brinksmeier and Oltmann Riemer designed the experiments;
Schnurbusch performed the experiments and materials investigation; Georg Schnurbusch and Oltmann Riemer
analyzed the data. Ekkard Brinksmeier contributed the experimental infrastructure; Georg Schnurbusch,
Ekkard Brinksmeier and Oltmann Riemer discussed the results and wrote the paper.
Conflicts of Interest: The authors declare no conflict of interest.

Abbreviations
a Hertzian contact circle radius LMdn average crack length (median)
ae depth of cut MRF magnetorheological finishing
cmax maximum crack depth ν Poisson’s ratio (workpiece)
cr radial crack length ν’ Poisson’s ratio (indenter)
E Young’s Modulus (workpiece) nW worktable speed
E’ Young’s Modulus (indenter) P indenter load
fL relative frequency R grain size (radius of the abrasive)
Fn normal force Sa average surface area roughness
H hardness SSD subsurface damage
K mismatch factor vc cutting speed

50
Inventions 2017, 2, 15

KIc fracture toughness vfr radial feed


L crack length WLI white light interferometer
Lcalc calculated crack length χr indentation constant
Lmax maximum crack length

References
1. Hed, P.P.; Edwards, D.F.; Davis, J.B. Subsurface damage in optical materials: Origin, measurement & removal.
In Proceedings of the Optical fabrication and testing workshop, Santa Clara, CA, USA, 2 November 1988.
2. Neauport, J.; Ambard, C.; Bercegol, H.; Cahuc, O.; Champreux, J.P.; Charles, J.L.; Cormont, P.; Darbois, N.;
Darnis, P.; Destribats, J.; et al. Optimizing fused silica polishing processes for 351 nm high power laser
application. In Laser-Inducted Damage in Optical Materials, Proceedings of the SPIE, Volume 7132, Boulder, CO,
USA, 22 September 2008; Exarhos, G.J., Ristau, D., Eds.; SPIE: Bellingham, WA, USA, 2008. [CrossRef]
3. Comley, P.; Morantz, P.; Shore, P.; Tonnellier, X. Grinding metre-scale mirror segments for the E-ELT ground
based telescope. CIRP Annals-Manufacturing Technology 2011, 60, 379–382. [CrossRef]
4. Miller, P.E.; Suratwala, T.I.; Wong, L.L.; Feit, M.D.; Menapace, J.A.; Davis, P.J.; Steele, R.A. The distribution
of subsurface damage in fused silica. In Laser-Inducted Damage in Optical Materials, Proceedings of
SPIE Volume 5991, Boulder, CO, USA, 19 September 2005; Exarhos, G.J., Guenther, A.H., Eds.; SPIE:
Bellingham, WA, USA, 2006. [CrossRef]
5. Lambropoulos, J.C. From Abrasive Size to Subsurface Damage in Grinding. TOPS 2000, 42, 17–18. [CrossRef]
6. Lucca, D.A.; Brinksmeier, E.; Goch, G. Progress in Assessing Surface and Subsurface Integrity. CIRP Ann.
1998, 47, 669–693. [CrossRef]
7. Wang, L.; Li, Y.; Han, J.; Xu, Q.; Guo, Y. Evaluating subsurface damage in optical glasses. JEOS:RP 6 2011, 6,
1–16. [CrossRef]
8. Menapace, J.A.; Davis, P.J.; Steele, W.A.; Wong, L.L.; Suratwala, T.I.; Miller, P.E. MRF Applications:
Measurement of process-dependent subsurface damage in optical materials using the MRF wedge technique.
In Laser-Inducted Damage in Optical Materials, Proceedings of SPIE Volume 5991, Boulder, CO, USA, 19 September
2005; Exarhos, G.J., Guenther, A.H., Eds.; SPIE: Bellingham, WA, USA, 2006. [CrossRef]
9. Menapace, J.A.; Davis, P.J.; Steele, W.A.; Wong, L.L.; Suratwala, T.I.; Miller, P.E. Utilization of
magnetorheological finishing as a diagnostic tool for investigating the three-dimensional structure of
fractures in fused silica. In Laser-Inducted Damage in Optical Materials, Proceedings of SPIE Volume 5991,
Boulder, CO, USA, 19 September 2005; Exarhos, G.J., Guenther, A.H., Eds.; SPIE: Bellingham, WA, USA, 2006.
[CrossRef]
10. Randi, J.A.; Lambropoulos, J.C.; Jacobs, S.D. Subsurface damage in some single crystalline optical materials.
Appl. Opt. 2005, 44, 2241–2249. [CrossRef] [PubMed]
11. Li, Y.; Zhenga, N.; Li, H.; Houa, J.; Lei, X.; Chena, X.; Yuana, Z.; Guoa, Z.; Wanga, J.; Guob, Y.; Xua, Q.
Morphology and distribution of subsurface damage in optical fused silica parts: Bound-abrasive grinding.
Appl. Surf. Sci. 2011, 257, 2066–2073. [CrossRef]
12. Klocke, F. Manufacturing Processes 2: Grinding, Honing, Lapping; Springer: Heidelberg, Germany, 2009;
pp. 251–269.
13. Wang, H.; Subhash, G.; Chandra, A. Characteristics of single-grit rotating scratch with a conical tool on pure
titanium. Wear 2001, 249, 566–581. [CrossRef]
14. Bollinger, L.D.; Gallatin, G.M.; Samuels, J.; Steinberg, G.; Zarowin, C.B. Rapid, non-contact optical figuring
of aspheric surfaces with Plasma Assisted Chemical Etching (PACE). In Advanced Optical Manufacturing
and Testing, Proceedings of SPIE 1333, San Diego, CA, USA, 1 July 1990; Sanger, G.M., Reid, P.B., Eds.; SPIE:
Bellingham, WA, USA, 1990. [CrossRef]
15. Suratwala, T.I.; Wong, L.L.; Miller, P.E.; Feit, M.D.; Menapace, J.A.; Steele, R.A.; Davis, P.J.; Walmer, D.
Subsurface mechanical damage distributions during grinding of fused silica. J. Non.-Cryst. Solids 2006, 352,
5601–5617. [CrossRef]
16. Huang, H.; Yin, L. High speed grinding performance and material removal mechanism of silicon nitride.
Initiat. Precis. Eng. Begin. Millenn. 2002, 416–420. [CrossRef]
17. Lawn, B.R.; Swain, M.V. Microfracture beneath point indentations in brittle solids. J. Mater. Sci. 1975, 10,
113–122. [CrossRef]

51
Inventions 2017, 2, 15

18. Lawn, B.R.; Evans, A.G.; Marshall, D.B. Elastic/plastic indentation damage in ceramics: The mediad/radial
crack system. J. Am. Ceram. Soc. 1980, 63, 574–581. [CrossRef]
19. Anstis, G.R.; Chantikul, P.; Lawn, B.R.; Marshall, D.B. A critical evaluation of indentation techniques for
measuring fracture toughness: I, direct crack measurements. J. Am. Ceram. Soc. 1981, 64, 533–538. [CrossRef]
20. Lawn, B.R. Partial cone crack formation in a brittle material loaded with a sliding spherical indenter. Proc. R
Soc. Lond. Ser. A 1967, 299, 307–316. [CrossRef]

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

52
inventions
Article
Stochastic Kinematic Process Model with
an Implemented Wear Model for High Feed
Dry Grinding
Michal Kuffa *, Fredy Kuster and Konrad Wegener
Institute of Machine Tools and Manufacturing, ETH Zürich, Leonhardstrasse 21, 8092 Zürich, Switzerland;
[email protected] (F.K.); [email protected] (K.W.)
* Correspondence: [email protected]; Tel.: +41-44-632-78-81

Received: 12 October 2017; Accepted: 14 November 2017; Published: 16 November 2017

Abstract: This paper considers heavy duty grinding with resin bonded corundum grinding wheels
and without lubrication and cooling. A vertical turning machine redesigned to a grinding machine
test bench with a power controlled grinding spindle is used in all of the experiments, allowing high
tangential table feed rates up to 480 m/min. This special test-rig emulates the railway grinding
usually done by a railway grinding train. The main test-rig components are presented and the
resulting kinematics of the experimental set-up is described. A stochastic kinematic grinding
model is presented. A wear model that is based on the kinematic description of the grinding
process is set up. Grain breakage is identified as the main wear phenomenon, initiated by grain
flattening and micro-splintering. The wear model is implemented into the stochastic kinematic
modelling. The workpiece material side flow and spring back are considered. The simulation results
are validated experimentally. The workpiece surface roughness is compared and a good agreement
between simulation and experiment can be found, where the deviation between the experiment
and the simulation is less than 15% for single-sided contact between the grinding wheel and the
workpiece. Higher deviations between simulation and experiment, up to 24%, for double-sided
contact is observed.

Keywords: wear modelling; self-sharpening; high-performance dry grinding; surface roughness

1. Introduction
Three important global trends of manufacturing technology in recent years were observed:
moving towards increasing performance in material removal rate, higher resulting surface quality,
and the reduction of coolant consumption. Chip transport from the cutting zone, cooling and
lubrication are the main tasks of the coolant. Environmental impact, costs, swarf, recycling,
and negative effects on the machine operator stand in contrast to the previously mentioned advantages.
Grinding is often used for finishing applications to generate smooth surfaces with high integrity
and low form deviations, but grinding is also yet increasingly challenging other manufacturing
technologies such as high-speed or high-performance milling and turning. The implementation of high
performance machining processes in general have become increasingly significant among cutting
processes and could already modify many industrial manufacturing process chains by substituting
one or more machining steps, and thus save cost.
As stated by Bhaduri and Chattopadhyay [1], but also by Kopac and Krajnik [2], high quality parts
require high-performance grinding technologies when considering not only superior surface finish and
high removal rate. The literature classifies high performance grinding not clearly distinguishable from
conventional grinding and covers different grinding techniques such as: High Speed Grinding (HSG),
High Efficiency Deep Grinding (HEDG), Speed-Stroke Grinding, High-Performance Surface Peel

Inventions 2017, 2, 31; doi:10.3390/inventions2040031 53 www.mdpi.com/journal/inventions


Inventions 2017, 2, 31

Grinding, High Work Speed Grinding, High Efficiency Grinding, Heavy Duty Grinding, and Creep
Feed Grinding (CFG) [2–11]. All of these grinding techniques are used with metalworking fluids
contrasting to the dry grinding process presented here.
Linke [12] gave in her review a broad overview on the recent grinding wheel wear examinations
and modelling and classified four main wear mechanisms: grain surface layer wear, grain splintering,
grain-bond-interface wear, and bond wear. Linke [12] distinguished between micro wear where cutting
edge sharpness is lost and macro wear as loss of tool profile.
Jiang et al. [13] introduced a three-dimensional (3D) model predicting the surface roughness
when considering wear and dressing effects indicating dressing as the most influential parameter.
A kinematic simulation predicting the surface roughness from Liu et al. [14] modelled three different
abrasive grain shapes and ductile and brittle fracture components by a single point diamond dressing
tool. Jiang et al. [13,14] judged dressing to be the most influential parameter on the predicted surface
roughness and grain shapes do not affect the resulting surface roughness. A stochastic process model
of abrasive wear, which is based on the Poisson distribution of the wear phenomenon, was presented
by Kacalak, Kasprzyk and Krzyżyński [15]. Meng and Ludema [16] presented 28 different erosion
wear equations from the literature having surveyed 5466 different papers. The synthesis of those
appears impossible to them.
In many dry grinding applications, resin bonded grinding wheels are used due to their
self-sharpening ability, cool cut, and damping capability. In this work, the high-performance
dry grinding process for surface face grinding with resin bonded corundum grinding cup wheels
is analysed at high feed speeds. Especially, the self-sharpening effect is analysed. An analytical wear
model considering grain failure phenomena is introduced.

2. Experimental Methods
The same test-rig as described in [17] is used. The following equipment is used for all experiments:
vertical lathe SEDIN 1525 with 20 kW grinding spindle typical in rail grinding, 3D-dynamometer Kistler
9366CC, roughness measuring system Taylor Hobson Talysurf, profile, and waviness measurement
sensor Micro Epsilon ILD 2300-10LL are used.
A detailed view of the experimental set-up is given in Figure 1. The marked tilt angle
is adjusted with a hydraulic cylinder and can be positioned in a positive and negative direction
within ±15◦ . The grinding wheel is clamped from outside with one fixed and two movable chuck
jaws. The grinding spindle is connected directly with the cover plate of the Kistler force measuring
platform. This guarantees the coordinate system conforms with force measurements without the
need of coordinate transformation. The axial movement of the grinding spindle and the adjustment
of depth of cut ap during the process are performed with the pneumatic cylinder mounted behind the
spindle. Two pillars guide the traverse with the spindle. The exact adjustment of the depth of cut is not
possible due to the power controlled spindle support that is used in railway grinding. Force control
is typical also for other high-performance dry grinding applications. The target current for the electric
engine is set and the pressure is continuously adjusted during the process in the pneumatic cylinder.
The tilt angle is adjusted with the hydraulic cylinder and the grinding spindle is guided on two milled
slides. The ring-like workpiece is placed on the rotary table that is fixed by four chuck jaws. The whole
grinding spindle construction is mounted in a frame. This frame guarantees the required stiffness and
strength for the process forces and torques. A laser sensor mounted on the right-hand column is used
for the surface measurements.
The cutting speed during all of the experiments is kept constant at 47 m/s at the outer
circumference of the grinding wheel and the feed rate of the workpiece is 120 m/min. A resin bonded
corundum grinding wheel with comparatively coarse mesh 20 is applied in the experiments. All of the
experiments are carried out without metalworking fluid. Workpiece material used is 58CrMoV4 with
hardness of 26 HRC. A 6 mm wide facet is ground in all of the experiments performed, however the
workpiece could be turned prior to any experiment to the desired width.

54
Inventions 2017, 2, 31

Figure 1. Experimental set-up: 1 grinding wheel, 2 force measurement system, 3 spindle motor,
4 vertical pneumatic cylinder, 5 frame, 6 milled slide, 7 workpiece, 8 laser sensor.

Tilting the whole grinding spindle around the y-axis, as shown in Figure 2, adjusts the tilt angle
while the approach angle of the grinding unit is connected to rotation around the x-axis.

(a) (b)

Figure 2. Schematic representation of the approach angle (a) and tilt angle (b).

Approach angle and tilt angle can be adjusted only in correlation to each other. A typical kinematic
interaction between the grinding wheel and the workpiece is a single-sided contact. For the tilt angle
−1.5◦ , the corresponding approach angle reaches zero leading to a dual-sided contact of the grinding
wheel. This fundamental kinematic difference depending on the tilt angle is shown in Figure 3.
Different surface patterns on the workpiece can be observed for both kinematic situations as shown
later. Additionally, the self-sharpening behaviour for the dual-sided contact is significantly lowered
due to the pressure being split on the two areas. Material removal rate and grinding wheel wear
change between both kinematic situations.

55
Inventions 2017, 2, 31

(a) (b)

Figure 3. One-sided contact (a) and double-sided contact; (b) of the grinding wheel after
initial self-sharpening.

3. Modelling

3.1. Stochastic Process Model


Based on the kinematic simulation presented in [17–19] a kinematic model with implemented
wear model is developed for resin bonded grinding wheels. The general scheme of the stochastic
process simulation with its main steps is shown in Figure 4.
The main input of the simulation is the grinding wheel geometry. Its generation is presented
and follows in six steps: the generation of the basic grain form, the modification of the grain shape,
the grain density definition, the extraction of the cutting profile, the positioning of the grain at the
grinding wheel surface, and the height distribution of the grain. The grinding wheel macro geometry
is defined by a circle formula. For abrasive grains modelled, five basic forms: cuboid, triangular prism,
two triangular pyramid, and dodecahedron are defined. These are re-shaper by edge and corner
fracture. This shape modification leads to a large amount of different abrasive grains corresponding
with the real grinding wheel, where crushed corundum abrasive grains are used. The abrasive grain
density is estimated experimentally leading and used as input for the grinding wheel generation.
For the abrasive grain height distribution, a triangular distribution delivers the most accurate results,
compared with uniform and normal distribution, and it is used in the grinding wheel topography
modelling. The resulting modelled grinding wheel topography is then compared with the topography
of the grinding wheel used experimentally by Abbott-Firestone curve. A good agreement can be found
as shown in Figure 5. The deviation towards the higher profile ratio is due to the modelled grinding
wheel, where only one layer of the abrasive grains is taken into account and the modelling concentrates
on the cutting edge space that comes into contact with the material. The deviation in the valleys of the
grinding wheel due to the approach of monolayer does not play any role for the process. The workpiece
geometry is meshed with nodes in radial and tangential direction. The nodes can be chosen arbitrarily.
In the meshed workpiece geometry, the resulting penetration between workpiece and grinding wheel
grains is saved as a surface roughness. The result quality is directly dependent on node quantity.
Process parameters that are used in the experiment and listed in the previous section are used
also in the simulation. The kinematic description used in the simulation is taken from [17].
After the definition of all the input parameters from Figure 4, including the grinding wheel,
the kinematic simulation can be performed. Abrasive grains, which cannot be in contact with the
workpiece, are deleted and only possible kinematic active grains are used for further calculation.
This allows, particularly for small cutting depths, a huge reduction of computational time.

56
Inventions 2017, 2, 31

Figure 4. Structure of the kinematic process simulation.

150
measured grinding wheel
100 modelled grinding wheel

50

-50

-100

-150
0 20 40 60 80 100
profile ratio [%]

Figure 5. Comparison of grinding wheel topographies for measured and modelled grinding wheel
by Abbott-Firestone curve.

57
Inventions 2017, 2, 31

In the simulation an ideal grinding wheel shape is loaded. According to the defined approach
angle, the geometry needs to be adapted. The dressed grinding wheel geometry is determined by the
contact points of this idealised grinding wheel with the workpiece profile. All of the cutting points
between the grinding wheel zero plane at outer and inner wheel diameter and between the workpiece
profile need to be found. The approach angle can be gained by contact point height difference at the
inner and outer grinding wheel diameter. In Figure 6, possible contact points are shown for grinding
process where the dressing angle α is defined.

Figure 6. Kinematic situation for double-sided contact between the grinding wheel and workpiece
with the connecting vectors.

The positioning vector of the contact point can be calculated in the workpiece and grinding
wheel coordinate system. Transforming both positioning vectors in the global coordinate system and
equalising them, an equation for each contact point can be defined. A cutting point on the grinding
wheel surface follows according to the equation:

R0P,SS = R0B + A0B · R BC + A0B · A BC · ( RCD + R DP ) (1)

where the SS subscript indicates that point P is the cutting point from a grinding wheel point of view.
R0P,SS is the connecting vector between an arbitrary point P at the grinding wheel surface and the
origin of the coordinate system indicated in Figure 6. R0B is the connecting vector from the origin
to the grinding wheel centre, RBC is the connecting vector from grinding wheel centre to the tilting
point, RCD is the connecting vector from the tilting point to the already tilted centre of the grinding
wheel, and RDP is the connecting vector from the tilted grinding wheel centre to the arbitrary point
on the grinding wheel surface P. The nomenclature of the vector is as follows: subscript describes the
connection between two points in a coordinate system. A0B and ABC are then transformation matrices
defined as: ⎛ ⎞
cos( β − β 0 ) − sin( β − β 0 ) 0
⎜ ⎟
A0B = ⎝ sin( β − β 0 ) cos( β − β 0 ) 0 ⎠ (2)
0 0 1
and ⎛ ⎞
cos ΩK sin Ω A sin ΩK cos Ω A sin ΩK
⎜ ⎟
A BC =⎝ 0 cos Ω A − sin Ω A ⎠ (3)
− sin ΩK sin Ω A cos ΩK cos Ω A cos ΩK

58
Inventions 2017, 2, 31

where ΩA is the approach angle and ΩK is the tilt angle. The transformation matrix A0B describes the
rotation of the local reference system B to the global coordinate system and ABC is responsible for the
adjustment of both, the tilt and approach angle.
The interaction between a pre-ground workpiece and a single grain is shown in Figure 7. From the
irregular cross section in cutting speed direction multiple overlaps with the cutting surface of the grain
might arise.

Figure 7. Intersection between one abrasive grain cutting profile and already ground workpiece surface
cross section seen in the direction of the cutting speed.

For every polygon, the cutting area can be calculated and summed up to a total cutting area
for each abrasive grain. This is used then as input for the force calculation. To calculate the force
for the whole grinding wheel, information about the force acting points for each grain is necessary.
It is assumed that this acting point is in the centre of gravity of the respective cutting area. The radial
position r and height position h of the centre of gravity are calculated with respect to the grinding
wheel referential plane as:

i −1
1
6 · At i∑
rt = · [r (i ) + r (i + 1)] · [(h(i ) − h(i + 1)) · (r (i ) · r (i − 1))] (4)
=1

i −1
1
ht = · ∑ [h(i ) + h(i + 1)] · [(h(i ) − h(i + 1)) · (r (i ) · r (i − 1))] (5)
6 · A t i =1

At is the partial polygon area from the intersection between the abrasive grain and the workpiece.
The centre of gravity can be then calculated from the partial polygon areas as:

1 i
A j∑
r= · rt · At (6)
=1

1 i
A j∑
h= · ht · At (7)
=1

The cutting force can be calculated according to an equation similar to the Kienzle equation as:

Fc = k c1,1 · Acut (t) (8)

where kc1,1 is the experimentally evaluated specific force and A represents the abrasive grain area
that is engaged with the workpiece and is orthogonal to the cutting direction. The force magnitude
delivered by the adapted Kienzle equation needs to be complemented by its direction. From the
3D abrasive grain geometry, as described in [17], the orthogonal cross-sectional cutting line is derived.
In addition to this, the projection of all other grain faces in contact with the workpiece on the
orthogonal cutting line is calculated. Each of these projected faces has its normal vector associated

59
Inventions 2017, 2, 31

with it. This allows the 3D abrasive grain geometry simplification without losing information.
Combining these projected normal vectors for one abrasive grain gives a resulting force direction
on one abrasive grain:
imax
v
F i = ∑ i · ni (9)
v
i =1 t

where imax is the total number of projected normal faces of one cutting grain, vi is the volume extracted
by the single face i, vt is the total volume extracted by the grain in the given time step, and ni is the
normal vector associated with the face i. This resulting force can be split into the cutting force Fc ,
showing in the direction of the cutting speed and the normal force FN being perpendicular to the
cutting force.
The resulting cutting force for the whole grinding wheel is then the sum of all the partial
forces acting:
n
Fc = ∑ Fci (10)
i =0

The normal forces are then calculated as:


Fc
FN = (11)
μ

where the friction coefficient μ = 0.3 is estimated from the experiment.


As a power controlled spindle is used in the experiment, the proper depth of cut needs
to be calculated. As explained above, the nominal current of the grinding spindle electric engine
is set and compared with the actually measured current. The measured current is a result of the
contact condition between the grinding wheel and the workpiece, and thus of the axial force on the
grinding wheel exerted by a pneumatic cylinder. This force needs to be counterbalanced by the
sum of the penetration forces of the individual active grains. With increasing force, the amount
of active abrasive grains and their penetration depth increases and the current rises. For each height
of the grinding wheel, the penetration depth of every single stochastically distributed, oriented,
and shaped active grain can be calculated, and thus the forces can be derived from the single grain
force model. Iteration is then performed until the sum of the individual grain forces equals the
externally applied axial force, pneumatic, and dead weight force, on the spindle. The resulting
experimentally measured vertical force is used in the simulation as the nominal value. An estimate
of the number of active abrasive grains is known, and thus the force can be distributed among them.
The initial grinding wheel position is defined as one where the highest abrasive grain touches the
workpiece surface. The simulation is performed with a stepwise increase of the depth of the cut by the
amount of one percent of an average abrasive grain diameter, as schematically shown in Figure 8.
If the sum of all vertical forces of the individual abrasive grains in contact with the workpiece is equal
to the nominal, experimental, force value, then the simulation continues at this depth of cut. As the
propagated wear of the abrasive grain can lead to an unbalance between the nominal and the real
vertical force, the depth of cut needs to be adjusted iteratively. In case an abrasive grain is completely
worn and is out of contact, the resulting forces are lowered and the depth of cut needs to be increased
again. The worn abrasive grain is in contact again and, additionally, new abrasive grains are potentially
cutting. The stepwise increase or decrease of the depth of cut is done until the balance between the
nominal and simulated force is found again. When the equilibrium between the simulated and
experimental forces is reached, the calculation of the workpiece surface roughness can start.
The workpiece material side flow and the workpiece spring back are modelled to determine the
ground profile. Based on Li [20], the side burr is modelled as a triangular ridge accompanying the scratch.
It is described by its base b and height h. Li [20] claims that the side flow is not significantly affected
by the abrasive grain rake angle and that the triangle base b is three times its height h which gives for h:

60
Inventions 2017, 2, 31

h = 1.27 · Acut 0.29 · v−


g
0.87
(12)

where vg is the grain velocity in m/s and Acut the area of the cutting profile being engaged with the
workpiece in m2 .

Figure 8. A stepwise adjustment of the depth of the cut in the simulation.

The material spring back model is based on a publication by Shaw and DeSalvo [21] about
hardness measurements. They claim that the material actively reacting on the imposed pressure on the
surface is maximum ten times the penetration of the indenter. The elastic material displacement can
be calculated as:
F·l
Δl = (13)
E·A
where Δl is the elastic material displacement, F is the normal force acting on the workpiece surface,
E is the elastic modulus, l is the material depth responding to the abrasive grain penetration,
and A is the workpiece area responding to the applied pressure. The force from experiment is in the
range from 1200 to 1400 N. This force is then divided between all of the active abrasive grains and
an average force is attributed to a single active grain. The length l is attributed to the workpiece
material amount, which answers to the pressure applied.

3.2. Wear Modelling


The main wear mechanism identified as responsible for the geometry change of the grinding
wheel is a fracture of the abrasives. This is a result of an initial abrasive grain flattening causing force
increase until fracture.
Grain fracture is observed experimentally, after reaching steady state, and is mainly in a vertical
direction—the direction of the normal force. The vertical component comes from the high normal
pressure due to the spindle weight and the power control. The horizontal component is resulting from
the cutting process and thus the spindle torque.
The main parameters influencing the grinding wheel wear are the cutting force, the cutting speed,
and the contact temperature. With increasing active time of the abrasive grain, the wear is propagating
until a critical load is reached. This behaviour can be expressed as:

dW − B
= K · Fc (t) · vc · e Tgrind (14)
dt
where dW/dt is the wear rate in mm/s, K is the wear factor, B is a constant coming from the Arrhenius
equation, where activating energy and universal gas constant are considered, Fc is the cutting force,
vc is the cutting speed for the abrasive grain, and Tgrind is the contact temperature between the abrasive
grain and the workpiece. Evaluation of the wear factor K is based on an experimental macroscopic
wear evaluation.
The cutting area Acut in Equation (8) changes with the wear progress W(t) and can be calculated as:

61
Inventions 2017, 2, 31

Acut (h(t)) = [(h0 − W (t)] · b (15)

where h0 is the initial abrasive grain height and b is the abrasive grain width assumed to be constant
during the whole cutting process as shown in Figure 9.

Figure 9. Abrasive grain geometry with parameters used for wear calculation.

The cutting force also depends on the cutting edge condition, which by wear becomes duller.
This leads to cutting force increase and can be considered by the introduction of a cutting edge dulling
factor M. The force can be then calculated as:
 
W (t)
Fc = k c1,1 · 1 + M · · Acut (h(t)) (16)
h0

With the cutting area being calculated according to Equation (15), the cutting force can
be calculated as:  
W 2 (t)
Fc = k c1,1 · b · h0 − W (t) + M · W (t) − M · (17)
h0
Substituting Equation (17) into Equation (14) leads to:
 
dW (− B ) W 2 (t)
= K · vc · e Tgrind · k c1,1 · b · h0 − W (t) + M · W (t) − M · (18)
dt h0

The whole term in front of the bracket is a constant in time and is renamed by H leading to:
 
dW W 2 (t)
= H · h0 − W ( t ) + M · W ( t ) − M · (19)
dt h0

This differential equation only for the wear and can be analytically solved to:

ec1 h0 M+c1 h0 + HMt+ Ht + 1


W ( t ) = h0 · (20)
ec1 h0 M+c1 h0 + HMt+ Ht − M

where c1 is a constant. As W(0) = 0 leads to:

ln(−1/h0 )
c1 = (21)
h0 · ( M + 1)

and thus to the solution:


e Ht( M+1) − 1
W ( t ) = h0 · (22)
e Ht( M+1) + M

62
Inventions 2017, 2, 31

The contact temperature between the abrasive grain and the workpiece defining H, given
by comparison between Equations (18) and (19) is calculated according to [22] as:
 0.53
2 · Rw · q f · α vw · l
Tgrind = 3.1 · · (23)
π · k T · vw 2·α

where Rw is the energy partition coefficient, qf is the heat flux, α is the workpiece thermal diffusivity,
kT is the thermal conductivity of the workpiece, vw is the workpiece feed rate, and l is the half length
of the band source.
A criterion needs to be established for the grain failure. Stresses acting on the abrasive grain are
estimated and compared to the critical stress value. The stress in the cutting direction can be estimated as:

Fc
σcut = (24)
Acut (h(t))

The stress in the normal direction can be calculated as:


FN
σN = (25)
A⊥

where FN is the normal force and A⊥ is the cross section of the abrasive grain projected in normal
direction. The normal force can be calculated from Equation (11).
Additionally, a shear stress can be calculated as:

Fc
τyz = (26)
b2
where Fc is the cutting force and b the width of abrasive grain.
For the calculation of the equivalent stress, the Rankine criterion is used:

σresult < σcritical = 300 MPa (27)

Plotting this resulting stress σresult qualitatively for different wear factors K and with M = 1
as simplifying assumption, different progressions can be observed as shown in Figure 10.

Figure 10. Qualitative representation of the resulting stress for a single abrasive grain for different
wear factors K.

63
Inventions 2017, 2, 31

From Figure 10 and together with the knowledge of the experimentally evaluated average time
for an abrasive grain failure and the calculated resulting stress, the wear factor can be evaluated.
Wear factor K = 508 × 10− 4 1/N is estimated as shown in Figure 11.

Figure 11. Resulting overall stress with the evaluated wear factor K.

With the wear factor K, the propagation of wear with time can be calculated according
to Equation (22).

4. Results and Discussion


In order to compare and validate the simulation with the experiment, the workpiece surface
roughness measurements are done. An indirect measuring method is carried out for all of the roughness
measurements. An imprint mass used for dental application is uniformly distributed on the workpiece
surface by dispense pistol, as shown in Figure 12. The overall imprint length allows for the maximum
measurement length of 80 mm. This imprinting method is time effective and allows for the conservation
of the current surface ground for later evaluation without losing the information. The accuracy of this
imprinting measuring method is proven in [23].

Figure 12. Workpiece surface imprint used for indirect roughness measurements.

All of the roughness measurements on the imprints are performed with a tactile roughness
measurement device Talysurf from Taylor Hobson, using diamond tip with 2 μm tip radius and
cut-off filter 2.5 mm. The measuring direction is along the feed rate direction, perpendicular to the
grinding grooves.

64
Inventions 2017, 2, 31

For all of the grinding experiments, the following process parameters were used: cutting speed
vc = 47 m/s, feed rate vw = 120 m/min, and grinding time 6 s. The grinding wheel was in steady
state. Pre-dressing, using the self-sharpening effect of the grinding wheel, was performed prior
to every experiment.
Figure 13 compares experimental and simulated results between the workpiece surface
topography for tilt angle 0.0◦ . Good agreement between the experimental and simulated roughness
values can be achieved. The roughness values Ra and the mean roughness value Rz are summarised
in Table 1. As validation of the model, the comparison of Ra and Rz might be used. For the purpose
of this research it is sufficient to verify the prediction of surface properties by Ra and Rz roughness
values. The workpiece surface anisotropy is in feed rate direction and in the cutting speed direction.
As the workpiece surface roughness, by railway grinding, in the feed rate direction is of interest,
Ra and Rz are sufficient.

(a) (b)

Figure 13. Comparison of surface topography for the experiment (a) and the simulation (b) using the
tilt angle 0.0◦ and the approach angle 0.12◦ .

Table 1. Average roughness Ra and average roughness depth Rz for experiment and simulation.

Tilt Angle Average Roughness Ra (μm) Average Roughness Depth Rz (μm)


Experiment Simulation Experiment Simulation
2.5◦ 9.8 9.9 61.8 59.4
0.0◦ 9.5 9.4 61.7 55.7
−1.5◦ 5.3 6.6 35.7 39.9

It can be seen, that for the tilt angle −1.5◦ in Figure 14, where the grinding wheel is in dual-sided
contact with the workpiece, the second part of the grinding wheel ruptures the roughness peaks.
The workpiece surface and the corresponding roughness values are lower as for the tilt angle 0.0◦ .
When considering the workpiece surface topology, a crisscross-like pattern can be observed for the tilt
angle −1.5◦ .
The mean roughness Ra reaches Ra = 6.6 μm for the simulation and Ra = 5.3 μm for the experiment.
The mean roughness depth Rz reaches Rz = 39.9 μm for the simulation and Rz = 35.7 μm for
the experiment.
When comparing the average Ra and Rz roughness values for tilt angles −1.5◦ , 0.0◦ ,
and 2.5◦ in Table 1, the simulation results at a tilt angle of 2.5◦ differ by maximum 12.2% for the
Ra values and by 14.4% for the Rz values. For the tilt angle 0.0◦ , the deviation by the Ra values
is maximum 5.2% and by the Rz values is 11%. For the double-sided grinding wheel contact, where
the tilt angle is −1.5◦ , the deviation for Ra is 24% and for Rz is 12%. This is caused by the associated
approach angle not being equal to zero. The self-sharpening effect of the resin bonded corundum
grinding wheel causes grinding wheel adaption on the workpiece geometry. Despite the relatively
large difference in the process kinematics for these both tilt angles, one size of the grinding wheel

65
Inventions 2017, 2, 31

cutting surface is always planar on the ground surface. This condition causes similar roughness
values for all different tilt approaches, except the one where the approach angle is zero (only for tilt
angle −1.5◦ ).

(a) (b)

Figure 14. Comparison of surface topography for the experiment (a) and the simulation (b) using the
tilt angle −1.5◦ and the approach angle 0.01◦ .

5. Conclusions
A kinematic process model with the focus on the wear modelling was presented. Time dependent
wear behaviour was presented. High feed rates were used for all of the experiments. The dependence
of surface roughness on the kinematic situation could be shown. With a dressed grinding wheel,
any adjusted tilt angle between ±15◦ shows the same workpiece roughness. This is because of the
self-sharpening of the resin bonded corundum grinding wheel and the adaption to the workpiece
geometry. The self-sharpening effect of the corundum grinding wheel shows that only the approach
angle 0.0◦ can improve the surface roughness, where a dual-sided grinding wheel contact is used.
An analytical approach for the wear modelling has been chosen based on previous studies.
The wear rate has been modelled as a function of the cutting force, the cutting speed, and the
temperature for every abrasive grain. The resulting exponential behaviour of the wear propagation
was used for subsequently determining the resulting stress on the abrasive grain, which is compared
to some critical stress level to determine the grain failure. With this model, the time delayed
failure of a grain could be described. Three wear parameters K, M, and B were introduced based
on experimental investigations. The average time for a grain failure was calculated and the wear factor
K = 508 × 104 1/N was estimated.
The simulation results show good agreement with the measured workpiece surface roughness.
When comparing the roughness values Ra and Rz, the deviation is by maximum 14.4% for tilt angle
2.5◦ , 11% for 0.0◦ and 24% for −1.5◦ .

Acknowledgments: The authors would like to thank the Commission for Technology and Innovation (CTI),
the Federal Office for the Environment (FOEN) and the Federal Office of Transport (FOT) for the fruitful
cooperation which made this work possible.
Author Contributions: Michal Kuffa conceived, designed and performed the experiments, analyzed the data and
wrote the manuscript. Fredy Kuster contributed to the analysis and the interpretation of results. Konrad Wegener
contributed to the modelling part, discussion and reviewed and approved the manuscript.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Bhaduri, D.; Chattopadhyay, A.K. Influence of Grinding Parameters and Substrate Bias Voltage in Dry
Surface Grinding with TiN-Coated Single Layer Galvanic cBN Wheel. Mater. Manuf. Process. 2011, 26,
982–990. [CrossRef]

66
Inventions 2017, 2, 31

2. Kopac, J.; Krajnik, P. High-performance grinding—A review. J. Mater. Process. Technol. 2006, 175, 278–284.
[CrossRef]
3. Schriefer, H. Kontinuierliches Wälzschleifen von Verzahnungen; Reishauer: Wallisellen, Switzerland, 2008.
4. Aurich, J.C.; Herzenstiel, P.; Sudermann, H.; Magg, T. High-performance dry grinding using a grinding
wheel with a defined grain pattern. CIRP Ann. 2008, 57, 357–362. [CrossRef]
5. Jackson, M.J.; Hitchiner, M.P. High Performance Grinding and Advanced Cutting Tools; Springer:
New York, NY, USA, 2013.
6. Rabiey, M.; Jochum, N.; Kuster, F. High performance grinding of zirconium oxide (ZrO2 ) using hybrid bond
diamond tools. CIRP Ann. 2013, 62, 343–346. [CrossRef]
7. Tawakoli, T.; Tavakkoli, S.J. High-Efficiency Deep Grinding (HEDG) of Inconel and Other Materials.
In Superabrasives; General Electric: Worthington, OH, USA, 1991; Volume 91, pp. D67–D82.
8. Chen, Z.Z.; Xu, J.H.; Ding, W.F.; Cheng, Z.; Fu, Y.C. High Speed Grinding of Nickel-Based Superalloy with
Single Diamond Grit. Adv. Mater. Res. 2011, 325, 140–146. [CrossRef]
9. Rowe, W.B.; Jin, T. Temperatures in high efficiency deep grinding (HEDG). CIRP Ann. 2001, 50, 205–208.
[CrossRef]
10. Hoffmeister, H.W. Hohe Zerspanungsleistungen Durch Schleifen Mit CD (Continuos Dressing) Sichere,
Werkstoffangepasste und Wirtschaftliche Prozessführung. Master’s Thesis, Technischen Universität
Carolo-Wilhelmina zu Braunschweig, Braunschweig, Germany, 1995.
11. Burkhard, G.; Rehsteiner, F.; Schumacher, B. High efficiency abrasive tool for honing. CIRP Ann. 2002, 51,
271–274. [CrossRef]
12. Linke, B.S. Review on Grinding Tool Wear With Regard to Sustainability. J. Manuf. Sci. Eng. 2015, 137, 060801.
[CrossRef]
13. Jiang, J.L.; Ge, P.Q.; Bi, W.B.; Zhang, L.; Wang, D.X.; Zhang, Y. 2D/3D ground surface topography modeling
considering dressing and wear effects in grinding process. Int. J. Mach. Tools Manuf. 2013, 74, 29–40.
[CrossRef]
14. Liu, Y.; Warkentin, A.; Bauer, R.; Gong, Y. Investigation of different grain shapes and dressing to predict
surface roughness in grinding using kinematic simulations. Precis. Eng. 2013, 37, 758–764. [CrossRef]
15. Kacalak, W.; Kasprzyk, M.; Krzyżyński, T. On Modelling of Stochastic Processes of Abrasive Wear and
Durability of Grinding Wheel. PAMM 2003, 2, 278–279. [CrossRef]
16. Meng, H.C.; Ludema, K.C. Wear models and predictive equations: Their form and content. Wear 1995,
181–183, 443–457. [CrossRef]
17. Kuffa, M.; Züger, S.; Kuster, F.; Wegener, K. A Kinematic Process Model and Investigation of Surface
Roughness for High Efficiency Dry Grinding. Procedia CIRP 2016, 46, 636–639. [CrossRef]
18. Vargas, G.E. Analyse und Simulation des Prozesses Honräumen von Gehärteten Innenprofilen
mit Diamantwerkzeugen. Ph.D. Thesis, Eidgenössische Technische Hochschule ETH Zürich,
Zürich, Switzerland, 2010.
19. Pinto Wagner, F. An Experimental and Numerical Approach to Investigate the Machining Performance
of Engineered Grinding Tools. Ph.D. Thesis, Eidgenössische Technische Hochschule ETH Zürich,
Zürich, Switzerland, 2008.
20. Li, X. Modeling and Simulation of Grinding Processes Based on a Virtual Wheel Model and Microscopic
Interaction Analysis. Ph.D. Thesis, Manufacturing Engineering, Worcester Polytechnic Institute,
Worcester, MA, USA, 2010.
21. Shaw, M.C.; DeSalvo, G.J. On the Plastic Flow Beneath a Blunt Axisymmetric Indenter. J. Eng. Ind. 1970, 92,
480–492. [CrossRef]

67
Inventions 2017, 2, 31

22. Kuffa, M.; Kuster, F.; Wegener, K. Comparison of lubrication conditions for grinding of mild steel with
electroplated cBN wheel. CIRP J. Manuf. Sci. Technol. 2016, 18, 53–59. [CrossRef]
23. Henerichs, M.; Egeter, M.; Liebrich, T.; Voss, R.; Wegener, K. Evaluation of the IWF-Wunder Reproduction
Method for Generating Positive Replica. Int. J. Autom. Technol. 2014, 8, 49–56. [CrossRef]

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

68
inventions
Article
Design Process Control for Improved Surface Finish
of Metal Additive Manufactured Parts of Complex
Build Geometry
Mikdam Jamal 1 and Michael N. Morgan 2, *
1 The Manufacturing Technology Centre Ltd., Ansty Park, Coventry CV7 9JU, UK;
[email protected]
2 Advanced Manufacturing Technology Research Laboratory (AMTREL), Faculty of Engineering and
Technology (FET), Liverpool John Moores University, Liverpool L3 3AF, UK
* Correspondence: [email protected]; Tel.: +44-(0)151-2312590

Received: 11 August 2017; Accepted: 6 December 2017; Published: 13 December 2017

Abstract: Metal additive manufacturing (AM) is increasingly used to create complex 3D components
at near net shape. However, the surface finish (SF) of the metal AM part is uneven, with surface
roughness being variable over the facets of the design. Standard post-processing methods such
as grinding and linishing often meet with major challenges in finishing parts of complex shape.
This paper reports on research that demonstrated that mass finishing (MF) processes are able to
deliver high-quality surface finishes (Ra and Sa) on AM-generated parts of a relatively complex
geometry (both internal features and external facets) under select conditions. Four processes were
studied in this work: stream finishing, high-energy (HE) centrifuge, drag finishing and disc finishing.
Optimisation of the drag finishing process was then studied using a structured design of experiments
(DOE). The effects of a range of finishing parameters were evaluated and optimal parameters and
conditions were determined. The study established that the proposed method can be successfully
applied in drag finishing to optimise the surface roughness in an industrial application and that it is
an economical way of obtaining the maximum amount of information in a short period of time with
a small number of tests. The study has also provided an important step in helping understand the
requirements of MF to deliver AM-generated parts to a target quality finish and cycle time.

Keywords: metal additive manufacturing; mass finishing; process optimization

1. Introduction
Presently, the surface finish of non-machined complex near net shape metal additive
manufacturing (AM) components is rough and uneven, having a relationship with build orientation,
machine build parameters, material and powder size. This is an increasingly challenging problem
as AM components are frequently of high complexity, rendering them even more difficult to finish,
which deleteriously impacts on post-processing time and labour, driving costs upwards. The relatively
poor surface finish is recognised as a barrier to wider adoption by industry and prevents fuller
exploitation of this disruptive technology. Mass finishing (MF) processes are often considered for
finishing AM parts, but it is challenging to know the appropriate abrasive media grade, size, shape,
process parameters, fixturing system and controls to deliver the desired ouptut.
A review of the literature shows that only limited scientific research has been published with
respect to mass finishing, more specifically associated with AM, and few mathematical or process
models exist. Most notably, Gillespie [1] is recognised as one of the earliest and major contributors
in this field, with his book being used as a guide to research. Further important publications [2,3]
are a reference for the scientific community and professionals alike. A more recent key paper [4] on

Inventions 2017, 2, 36; doi:10.3390/inventions2040036 69 www.mdpi.com/journal/inventions


Inventions 2017, 2, 36

topographical development and areal characterization provides guidance on the optimal parameters
and sampling method to characterise this surface type for a given application. A number of trade
articles and empirical studies have also been published [5–8]. These describe, in general terms, how
surface finish and material removal are controlled, but these have only limited usefulness as they
provide little insight into the comparative performance of different tool materials; they fail to explain
fully and scientifically the relationships between key process parameters.
It is reasoned that, with the benefit of increased data, the relationship between finishing parameters
and surface finish could be predicted more accurately, thus leading to more robust design tools
that could serve to improve quality in the AM process and the final component. The introduction
of this design aims to advance both AM and surface engineering technologies, with the ability to
enhance AM component surface finish quality through post-processing methodologies to suit the
AM parts. Post-processing costs and AM manufacturing costs will thereby be reduced, increasing the
competiveness of the AM manufacturer and the surface finishing company and leading to increased
profitability. Achieving surface finish specification will allow other industries to increase their uptake
of metal AM, utilise increased part complexity in their processes, and take advantage of open surface
coating opportunities for these parts.
The aims of the study were to obtain a quantitative understanding of the capability and efficacy
of MF processes in delivering an AM part with a target surface finish over a range of system input
conditions and to create a dataset and inferred rules based on process informatics. This would serve as
a generic platform to aid the user with selecting optimal production parameters for the generation of
parts with a specified output surface finish and/or production cost.

2. Mass Finishing Process Review


There are a number of different MF processes in common use in industry. Among these are
barrel, centrifugal and vibratory. Vibratory systems have become the dominant technique due to
the advantages inherent to the method in terms of ease of use and material handling. There is no
fixturing in these systems, so the parts can flow freely with the media. Such systems fall into two broad
categories in terms of the equipment being used: round bowl and rectangular tub designs [9]. These
systems are typically modest in size and used for the deburring and finish processing of smaller
components. The capability to process larger batch lots of various sized parts is important to the
competitiveness of such processes. The mass finishing processes are described in the following sections.

2.1. Mass Finishing Processes

2.1.1. Centrifugal Disc


Centrifugal disc machines were designed to combine the three-dimensional action advantage of
a vibratory bowl with the rapid cycle time of the centrifugal barrel. Figure 1 shows an image of the
principle of a centrifugal disc machine. The centrifugal force and the rotating action of the disc sends
the mass outwards towards the chamber wall. The mass then slows down and returns to the centre to
complete the cycle [10]. Centrifugal disc machines are also used for their fast processing and cycle time
(typically 10 to 30 min), their capability for batch production and their control in process [1]. However,
due to the small capacity of the chamber, the workpiece should not exceed 30 cm in length.

2.1.2. High-Energy Centrifugal Barrel


Centrifugal barrel machines often provide finishing results that cannot be achieved in a standard
vibratory process. A centrifugal barrel machine is typically comprised of four barrels horizontally
mounted between two main drive turrets, as is shown in Figure 2. As the two main turrets rotate,
all barrels rotate in the opposite direction of the turret. The capacities of centrifugal barrel machines
range from 0.5 cf to 12 cf of processing volume.

70
Inventions 2017, 2, 36

Figure 1. Centrifugal disc machine principle (LJMU).

Figure 2. High-energy centrifugal barrel machine principle showing media and parts in each barrel.

This process requires manual loading and unloading as well as parts and media separation.
Processes ranging from heavy grinding to fine polishing can be accomplished with centrifugal barrel
finishing. Parts can be processed dry in a few select media types in a centrifugal barrel. Both wet or
dry, a great deal of heat and pressure are generated within each barrel during this method. Heat must
be considered from both a safety and finishing standpoint when centrifugal barrels are utilised.

2.1.3. Fixture Mass Finishing


The purpose of a fixturing method is to hold and give positional and/or rotational variation to
the part and its interaction with the media. In the case of a fixed part, immersed in media and rotated
at speed (e.g., drag finishing), the part edges and surfaces interact with loose media and a higher force
is generated than that developed in conventional mass finishing processes, in which the part is placed
randomly in the machine chamber and depends on the media motion to achieve the surface finishing
result. Fixture arrangements also promote quicker cycle times. Furthermore, a benefit of the fixturing
arrangement is that contact between parts is prevented. The best examples of fixture processes are drag
finishing and stream finishing, which give high-quality results with short processing time and high
reliability [11]. In this study, automation and optimisation of fixture finishing process were employed
to help minimise human intervention.

Drag Finishing
Drag finishing machine technology is used for high-quality and sensitive workpieces that need
to be ground or polished without contact between parts; it is available in dry and wet processing.

71
Inventions 2017, 2, 36

Workpieces are fixed in a special arrangement of workpiece holders. Holders are dragged in circular
motion through the container that is filled with the loose abrasive or polishing media. The main
spindle is also rotated at different speeds, either clockwise or anticlockwise. Figure 3 shows an image
of the drag finishing machine principle.

Figure 3. Drag finishing machine principle.

The mass finishing process is significantly dependent on pressure and speed. In the drag finishing
process, the interaction between the high-speed workpiece and media will create high-pressure contacts,
leading in a short time to the development of a high-quality surface finish and precise rounding of
the part edges, which can otherwise only be obtained by the hand-polishing method. By controlling
variables such as immersion depth and rotary speed, it is possible to reach rates of material removal
up to 40 times higher than in conventional (vibratory, centrifuge) mass finishing and about 4–5 times
faster than high-energy vibratory systems. In addition, the existence of a double drive system with
two electric motors allows workpiece edges to be rounded uniformly. This also enhances the tool life.
The applications for drag finishing range from aggressive deburring and edge breaking through to
high-gloss polishing.

Stream Finishing
Stream finishing is a new method of mass finishing. It is also termed abrasion finishing, though it
is not always used as an abrasion process but also for lapping and polishing, depending on the output
target. It is able to finish workpieces of length up to 200 mm and weight of 20 kg, including the holder.
Figure 4 shows an image of the principle of a stream finishing machine. The workpiece holder
is rotated either clockwise or anti clockwise to allow the workpiece itself to align to the direction of
stream flow media. The media flow is generated through the bowl rotation at a specific speed. This
configuration helps reduce the effect of transverse forces and lowers the risk of the tool breaking or
becoming deformed [1]. Stream finishing is short in processing time, simple in operation, and has
excellent reliability. However, despite the high processing force it can generate a fine surface roughness
value (Ra) less than 0.03 μm.
The machine shown is able to process six workpieces in unison in a single holder fitted with
a quick release mechanism to ensure that the workpiece is finished over the entire surface. Workpiece
holders (examples: jaw chuck, drill chuck, collets and gripper) are available possessing different
independent features that can rotate around their own axis at the same time. The processing time can
be reduced significantly by adapting high-speed rotating workpiece holders (up to 2000 rpm).

72
Inventions 2017, 2, 36

Figure 4. Stream finishing machine principle (LJMU).

2.2. Mass Finishing Media


There are five major materials of loose abrasive media (ceramic, plastic and synthetic, metal,
precision, and micro-bite). The internal composition of conventional media is the source of the cutting
action, weight, wear, and level of surface finish, all of which define the capabilities of the media.

2.2.1. Shape, Size and Weight


There are approximately 15 different major shapes of loose abrasive media led by triangular,
conical, cylindrical, spherical and tetrahedral. Each abrasive media is used for a specific application
depending on the workpiece configuration. Shape is an important factor in media selection. The major
considerations are that the shape of the loose abrasive media chosen should permit access to all
surfaces of the workpiece—meaning any ledges, holes, or other geometric features—for the deburring
or finishing operation, and it must allow an easy separation of the media from the workpiece treated
at the end of the cycle. In some cases, a mixture of media shapes may be most effective.
The size of loose abrasive media is important for several reasons. It helps to keep the workpieces
separate from each other. The right choice can allow the media to reach all surfaces to finish or deburr.
Larger media have a faster cutting rate and produce smoother surfaces on workpieces. However,
the use of small media in high-energy methods of mass finishing can improve the surface finish
without increasing the cycle time. The weight of the loose abrasive media is another important factor
in media selection.
According to [1], heavier media allow faster cutting. However, heavier media apply more pressure
than lighter media, which can cause thin or ductile workpieces to bend or distort.

2.2.2. Media Capability


The capability is the ability of the media to deburr corners, edges and other surface requirements
with uniformity (constant performance over time). The capability of a media is principally defined by:
the machining parameters; the media wear rate; media composition (cutting action, weight, wear and
surface finish); the ability to finish a range of workpieces (i.e., shape, size, material); and quality of the
finishing. Hard media will work longer, but softer media will cut faster. In terms of media wear, most
plastic, abrasive, or ceramic media get smaller over time, which can affect the finishing operation if
they are lodged in holes and other recesses, which requires removing them manually.
The efficiency of material removal is strongly determined by the sharpness, hardness and density
of the abrasive cutting edges. A media with an aggressive cutting surface will result in a poor surface
finish and a media with a fine cutting efficiency will result in a good surface finish. This aspect is
essential to understand the capability of the media assessed in this study.
The abrasive grain distribution on the conical shape media has been shown [1] to be largely
independent to stock removed. A study directed by Moore [12] investigating the relationship between
the volume of workpiece abrasive wear and the mean diameter of the abrasive particles shown that

73
Inventions 2017, 2, 36

it is possible to find the most economical solution by determining the ratio of media to workload.
The results of that study indicate that any change in the optimum ratio can produce a significant
effect on the quality of finishing achieved [1]. It can also be noted that, for the range of materials
used in the study, increasing the diameter of the abrasive grain increases the volume that is removed
from the workpiece. Domblesk and Dennis [13,14] suggested that the particle shape is not the only
variable that controls the behaviour of manufactured abrasives. The relationship between particle
shape and packing behaviour needs to be understood before the former can be used to accurately
predict the wear-rate.

2.3. Additive Manufacturing (AM) Artefact Design for Finishing

2.3.1. Design Constraints of Part to Be Finished


The geometry for the finishing artefact is influenced by three main factors:

­ Design space: The artefact should be bound within a Ø40 mm × 80 mm design space. This was
required so that the artefact could move freely within the high-energy machine, as shown in
Figure 5.
­ Selective Laser Melting (SLM) using Renishaw AM250. Design points to take into account:

• Build direction
• Minimum feature size (refer to design features)
• Build supports
• Size of part (SLM machine can build parts larger than this design space)
­ Design features: The part had to include a variety of design features within a compact design
space. Including: thru holes, free-form surfaces and sharp corners.

Figure 5. High-energy centrifuge and design space for part (grey shaded area).

2.3.2. Artefact Design Summary


­ Part Size: 25 mm × 25 mm × 80 mm
­ Design Features (Figure 6)

• Free-form surface
• 2 × Angled surfaces (45 degrees to horizontal)
• Concave and convex features
• 2 × cut out sections (5 mm × 5 mm × 25 mm)
• 10 mm and 15 mm thru holes
• 3 × blocks (1, 2 and 3 mm height)
• Flat surface (25 mm × 65 mm)

74
Inventions 2017, 2, 36

­ Fixturing point:

• 2 mm pilot hole for internal thread.

Figure 6. The resultant additive manufactured (AM) artefact design features.

2.4. AM Build Strategy


Build height: 32.5 mm
Build time: ~40 h
Supports: Minimal supports are required for this build
Machine: Renishaw AM250
Parts per build: Maximum of 10 parts per build
The base materials used in this study were Ti-6Al-4V. All parts were built using a Renishaw AM250
selective laser melting (SLM) machine. According to a complex part manufacturing setting, the vertical
strategy is used to create a Ti thin section sample with layer thickness of 30 μm. The Ti-6Al-4V powder
diameters used for experiments were located within the range of 10–20 μm and the composition
conformed to the standard, Table 1. Argon gas is used during the process in order to protect the
oxidation of surfaces. The gas is supplied through the powder projection nozzle of the additive
manufacturing machine. The operating parameters used for the build samples are listed in Table 2.
The average surface roughness of the base material is around 20 μm. The build configuration on the
plate is shown in Figure 7.

Table 1. Chemical compositions for Ti-6Al-4V used in the study (wt %).

Al V O N C H Fe Ti
5.5–6.75 3.5–4.5 0.2 0.05 0.08 0.015 0.3 Bal.

Table 2. Build parameters.

Build Parameter Value


Power (W) 300
Beam size (μm) 85
Point distance (μm) 60
Effective speed (mm/s) 500
Hatch spacing (μm) 50

75
Inventions 2017, 2, 36

Figure 7. AM building strategy (showing eight parts).

2.5. Experimental Design and Setup


The experimental analyses were carried out on the comparison between different mass finishing
processes of drag, stream, disc centrifuge and high-energy centrifuge. The AM metal parts in Stainless
Steel 316 were used in this investigation with average surface roughness values, following build and
sand blasting, of 1.7 μm (±0.15 μm). The liquid compound SC15 was used for the tests. The proposed
techniques were compared according to the 2D surface finish parameters (Ra, Rz, and Rt) and areal
surface finish parameters (Sa, Sz, and St) achievable at constant processing time intervals of 30, 60, 90
and 120 min. The experimental parameters are summarised in Table 3.
Three different media were used during the finishing process: plastic (OTEC DS, conical), ceramic
(OTEC KM10 and KXMA24, tetrahedral) and Al Oxide (spherical). The experimental tests that involved
drag finishing were performed by setting the head speed at 50 rpm, and varying the spindle speed over
the values of 20, 40, and 80 rpm. In the case of stream finishing, the experimental tests were performed
by setting the spindle speed at 30 rpm, and changing bowl speed over the values of 20, 40 and 60 rpm.
The position of the workpiece in the bowl was selected carefully at the maximum stream flow region.
In both cases full immersion of the workpieces were employed to ensure better performance results.
The work holding design has been fabricated according to the workpiece geometry. The angled head
was selected to be zero in all operation processes in order to ensure optimal contact between the media
and workpiece.
The experimental tests relating to the centrifugal disc finishing, were performed by varying the
rotating speed within the values of 150, 200 and 250 rpm, while the rotation speed kept at 200 rpm for
the high-energy centrifuge.
The 2D surface roughness of the workpieces was inspected by tracing the surface using a Talysurf
surface instrument at several positions along the surface before and after finishing processes (stress
force = 0.75 mN; standard cut-off = 0.1 mm; transverse length = 0.8 mm; amplitude height 2.5 μm;
stylus speed = 0.5 mm/s) Three surface roughness measurements were performed on the workpiece
surface at each time interval; the average of the group results are presented in this study.
A three-dimensional optical profiler instrument was used to assess the morphology of the
specimen before and after the machining process. The optical profile instrument (GFM) uses the

76
Inventions 2017, 2, 36

phase measurement fringe projection technology method and is a good tool for characterising surface
height variations and quantifying the amplitude of peaks and valleys (roughness profile) resulting
from the machining process.

Table 3. Full range of machining parameters. HE: high-energy.

Drag Finishing
Main head speed (rpm) 50
Spindle speed (rpm) 20, 40, 80
Processing time (min) 30, 60, 90, 120
Immersion depth (mm) 300/full immersion depth
Angled head (degree) 0
Media Plastic KM10, Ceramic KXMA24
Compound SC15
Dosing (%) 3.9
Stream Finishing
Spindle speed (rpm) 30
Bowl speed (rpm) 20, 40, 60
Processing time (min) 30, 60, 90, 120
Immersion depth (mm) 300/full immersion depth
Angled head (degree) 0
Media Plastic KM10, Ceramic KXMA24
Compound SC15
Dosing (%) 3.9
Centrifugal Finishing
Disc speed (rpm) 150, 200, 250
Processing time (min) 30, 60, 90, 120
Media Ceramic KXMA24
Compound SC15
Dosing (%) 1
HE Centrifugal
Finishing 150, 200, 250
Disc speed (rpm) 30, 60, 90, 120
Processing time (min) Ceramic
Media SC15
Compound Wet condition
Dosing (%) 1

3. Results
The results of the comparative evaluation between the four advanced finishing process techniques
(drag, stream, disc, and high-energy (HE) centrifuge) are summarised in Figures 8–11, which provide
the surface roughness values at different stages of the manufacturing process.
The results show that each of the finishing processes improved the starting workpiece morphology
by progressively cutting the peaks and obtaining a smooth surface based on the setting of the
operational parameters. The surface roughness profile (Figures 8 and 9) shows that the valley surfaces
remain largely unaffected by the finishing process, this may be attributed to the media size, which
cannot reach these regions except in stream finishing where fine spherical Al Oxide media was
employed. It is clearly noticed, the relatively high spindle head speed of drag and stream, reduces
the average surface roughness over the processing time regardless of the abrasive tool. However,
the results demonstrated that the rate at which surface roughness decreases over the maximum cycle
time is related to the AM part orientation. Table 4 shows the decreases in surface roughness using
various mass finishing processes at maximum processing time of 120 min.

77
Inventions 2017, 2, 36

Table 4. Surface roughness values (average) of mass finishing processes at maximum processing time.

Finishing Process Top Surface Ra (μm) Bottom Surface Ra (μm) Side Surface Ra (μm) Angle Surface Ra (μm)
Drag 0.28 0.33 0.36 0.63
Stream 0.29 0.27 0.24 0.36
Disc 0.47 0.21 0.76 0.93
HE Centrifuge 0.48 0.39 0.42 0.75

Figure 8. Surface roughness comparison on bottom surface (blue arrows show direction of measurement).

Figure 9. Surface roughness comparison on top surface (blue arrows show direction of measurement).

Figure 10. Surface roughness comparison on angle surface (blue arrows show direction of measurement).

78
Inventions 2017, 2, 36

Figure 11. Surface roughness comparison on side surface (blue arrows show direction of measurement).

Mass finishing processes of the external surfaces, and easily accessible internal surfaces, to specification
can be accomplished by traditional processes. However, most of these processes are inflexible and provide
little control, nor can they access internal surfaces of complex shapes. For these type of challenges stream
finishing can be a viable solution. Figure 12 shows the surface roughness (Ra) values of the internal hole
surface as a function of processing time using the stream finishing process and Al Oxide media. It is clear
that the surface roughness decreases to the value of 0.41 μm at the maximum processing time of 180 min.

Figure 12. Stream finishing internal surface roughness values as a function of processing time (blue
arrows show direction of measurement).

Figures 13 and 14 show the comparison of areal surface roughness (Sa, Sz, and St) of the side
surface at maximum processing time of 120 min using different finishing technologies and media
grade (ceramic media have been used for disc, HE centrifuge, and drag finishing processes and Al
Oxide media for the stream finishing process). Figure 13 shows that the decreases in surface roughness
Sa of the side surface using drag, stream, disc, and HE centrifuge at maximum processing time of
120 min were 3.2, 1.7, 2.3, and 1.3 μm, respectively. Similarly, in Figure 14 the HE centrifuge achieved
the lowest Sz, and St values followed by stream, centrifuge and drag finishing processes.
Figure 15 shows an example of the 3D mapping assessment of AM part morphology using stream
and HE centrifuge finishing processes. The 3D maps were obtained before and after 120 min of finishing.
The results demonstrated that, after 180 min, the built layer peaks of the workpieces are very thin
when processed with stream finishing and completely disappear with the HE centrifuge. This suggests
that the HE centrifuge finishing process with ceramic media has the most aggressive uncontrolled
effect on the AM part, which may cause over-finishing of the workpiece surface in the context of poor

79
Inventions 2017, 2, 36

edge/radius retention, feature losses and breaching of tolerance criteria. This over-finishing may cause
a failure of the workpiece due to the high surface erosion and may force a re-design.

Figure 13. Sa comparison at maximum processing time of various mass finishing processes.

Figure 14. Sz and St comparison at maximum processing time of various mass finishing processes.

(a)

Figure 15. Cont.

80
Inventions 2017, 2, 36

(b)

(c)

Figure 15. A 3D map assessment of AM part morphology before and after finishing processes.
(a) Pre-processing 3D areal surface roughness parameters; (b) Stream finishing 3D areal surface
roughness parameters; (c) HE centrifuge finishing 3D areal surface roughness parameters.

Despite the wide versatility of mass finishing processes, some potential process limitations should
be noted. For the non-controllable finishing processes (disc and HE centrifuge), it can be difficult to
selectively treat certain areas of the part to the exclusion of other areas, which might have critical
dimensional tolerance requirements. Unless masked or fixtured, all exterior areas of the part will be
affected by the process to a greater or lesser degree, with effects on corners and edges being more
pronounced than those on flat areas, and with interior holes, channels, and recesses being relatively
unaffected in the more common processes. However, care must be exercised in media size, shape
selection, machining parameters and maintenance to prevent media lodging in holes and recesses,
which might require labour-intensive manual removal and prohibit further abrasive/polishing action.
Some parts have shapes, sizes, or weights that may prevent them from being finished in some
traditional mass finishing processes because of the risk of impingement from part-on-part contact.
However, drag and stream finishing processes can be a viable solution for such challenges.

4. Process Optimisation System of Drag Finishing


For this study, the core of the process optimisation system was based on readily available
and well-understood proprietary statistical and mathematical software packages. The complete
optimisation tool used ANOVA, the Taguchi experimental methodology and the response surface
methodology (RSM) coupled with a MATLAB—Graphical User Interface (GUI). These tools were used
to develop a surface roughness prediction model of advanced mass finishing technology under various

81
Inventions 2017, 2, 36

machining parameters conditions. The model can then be used to estimate process parameters needed
to achieve a desired roughness of a workpiece.
The usefulness of the system is not restricted to optimisation in the production context. It can
also be used as design tool or as a production planning aid. As a design tool it can provide optimal
parameter sets for materials new to the process or not in common use, and as a planning tool to
plan production cycles to match related operations and to avoid misalignment of outputs with other
processes, or to ensure maximum machine utilisation or indeed both. Further, it has potential to be of
great benefit to machine tool salespersons and production engineers in the field.

4.1. The Taguchi Method


Taguchi is a statistical experimental tool designed to calculate a single or multiple groups of data
(variable) to find the optimal and regularity in the operating environment (response). It is also used to
identify the controllable and un-controllable factors in order to minimise the effect of noise factors.
The orthogonal array (L8) matrix is designed to investigate the effect of each factor independently
from others in order to reduce time and cost. The main objective is to identify the ranks of each factor
that affect the mean response.
The statistical evaluation of performance required use of a signal-to-noise (S/N) ratio approach to
measure deviation of the quality characteristic from the design value.
The S/N ratio can be used instead of the average value for the evaluation characteristic in the
optimum variable analysis [15–17].
The S/N ratio approach can be divided into three quality tools: “the smaller the better”, “the nominal
the better”, and “the higher the better” when characterizing the response quality for engineering
processes. In this study, “the smaller the better” surface roughness was considered in investigating the
influence of machining variables in the drag finishing process. The S/N ratio of “the smaller the better”
quality characteristics can be defined by Equation (1):

S n Y2
N
= −10 log ∑ i =1 n
, (1)

where (Y2 )/n is the mean of the square of measured data at i test and n is the number of repetitions.

Experimental Design and Setup


Experimental investigations were carried out using a drag OTEC machine. AM parts were used
in the experiments with average surface roughness values of 1.7 μm (±0.15 μm). Liquid compound
SC15 at a constant dosing rate of 3%. Ceramic and plastic media KM10 were employed in the tests.
All experiments were performed at a constant processing time of 90 min.
The machining parameters were selected as follows: the experimental tests were performed with
variable spindle speed, DSS within range (10 ≤ DSS ≤ 80) rpm, the head speed, DHS various within
range (10 ≤ DHS ≤ 60).
Another important parameter that should be considered in the drag is the angle of the head
holding the workpiece. Due to the shape of the AM part, the angled head was selected to be zero
(vertical position) in all tests using a special holding arrangement to ensure continuous flow of media
on the workpiece surface during operation.
Areal Sa surface roughness tests were conducted using a three-dimensional optical profiler
instrument to investigate the morphology of the specimen before and after the machining process.
Three surface roughness measurements were performed on each workpiece material, the average
of the group results is presented in this study. However, this work can be extended to include the
process optimisation system of Taguchi, ANOVA, and surface response methodologies to develop
a mathematical model used to predict the surface roughness of the AM component, as shown in
Figure 16. In general, the process optimisation system is based on 10 steps of planning, performing

82
Inventions 2017, 2, 36

and evaluating results of a matrix of experiments to determine the performance levels of control
parameters [18].

Figure 16. A schema of the process optimisation system.

4.2. Results

Analysis of the Signal-to-Noise (S/N) Ratio (Taguchi Method)


The S/N number represents the ratio between desirable and non-desirable values. The process
parameter with the highest value has the most significant effect on the optimisation process [15]. In this
study, the Taguchi optimisation process was selected with six degrees of freedom, three factors and
two levels. The process parameters and their levels in the drag process are presented in Table 5.

Table 5. DOE layout and drag finishing results using Taguchi method.

Process Parameter Level 1 Level 2


Spindle speed (DSS) 20 80
Head speed(DHS) 20 50
Media type Ceramic Plastic

The process variables and their levels were set over a wide range of parameters in order to
cover a comprehensive performance capability of the finishing processes. A set of experiments were
conducted using the dag finishing process and the L8 orthogonal array with eight experiments. Table 6
shows the DOE layout and results from the drag finishing processes.
Due to different units within the process parameters, each factor was coded with two levels.
The influence of the interaction between variables is neglected in the Taguchi method. Regardless of

83
Inventions 2017, 2, 36

the machining process characteristics, the greater the S/N value, the better the performance. Based on
the S/N ratio analysis, the optimal drag finisher performance for best surface roughness was obtained
at 80 rpm spindle speed (level 2), 50 rpm head speed (level 2), and Ceramic media (level 1).
Optimal finishing conditions for the drag process are shown in Table 7. Table 7 indicates that the
strongest effect on the finishing process is given as rank 1 and is associated with the media, followed
by rank 2 (the head speed), while rank 3 (the spindle speed) has the weakest effect on the process.

Table 6. Drag finishing process parameters and their levels.

No. DSS DHS Media Sa (μm) S/N


1 1 1 1 2.0 −6.020
2 1 1 2 4.3 −12.669
3 1 2 1 1.9 −3.521
4 1 2 2 5.2 −10.370
5 2 1 1 2.5 −5.575
6 2 1 2 7.9 −14.320
7 2 2 1 1.5 −7.958
8 2 2 2 3.3 −17.952

Table 7. S/N ratios for surface roughness using drag finisher.

Level DSS DHS Media


1 −9.646 −11.150 −5.769
2 −9.951 −8.447 13.828
Delta 0.305 2.704 8.509
Rank 3 2 1

Figure 17 shows the principal variations caused by changes to the S/N ratio. The strongest
influence of the factor on the finishing process is measured by differences values. The factor that
contributes to the higher difference in the mean S/N ratio over the selected design control factor,
the more significant the impact. Optimal finishing conditions of design control factors of drag finishing
process can then be determined from the S/N response graphs. As displayed, the media type shows
the higher slope of the S/N ratio, which identifies the significant impact on the finishing process.
The results show that the surface roughness decreases with the increasing drag head speed. The S/N
ratios also identify that spindle speed had the least influence of the three factors.

Figure 17. Effect of drag process parameters on surface roughness.

84
Inventions 2017, 2, 36

5. Conclusions
A comparison of results obtained using spindle finishing (drag and stream), centrifugal disc
finishing and HE centrifuge processes using conventional abrasive granulate (Al Oxide and ceramic)
media on the surface finish of an AM component has been undertaken. The results for the cases studied
suggest that the stream finishing process is most efficient in reducing the surface roughness followed
by HE centrifuge and drag finishing, then finally the centrifugal disc finishing process. However, it is
to be noted that the media used in the drag machine was not the optimal type for this study.
The abrasion potential of the drag and stream finishing processes is high, especially when operated
at the fastest spindle speed of 80 rpm and bowl speed 60 rpm, respectively. This may be attributed to
the different hydrodynamic behaviour of the abrasive media in the processing medium. The impacts
with abrasive media take place at high speed, which is generated from the difference between the
workpiece rotating speed and the quasi-static abrasive media for the drag finishing process and from
the difference between the workpiece rotation speed and abrasive media rotation flow for the stream
finishing process. In the centrifugal finishing process, the workpiece is immersed into the abrasive
media and travels with the flow.
Despite the high rotation speed of the disc in the centrifugal finishing process (250 rpm),
the relative speeds between the abrasive media and the workpiece are expected to be lower because it
only comes from the resultant difference in the system force acting on the abrasive medium. This speed
is much higher in the HE centrifuge process, resulting in high-energy contact ,which alters the AM
surface texture substantially, not only reducing surface peaks but generating a surface in which the
positioning of the peaks has been altered appreciably.
However, it is the impact speed together with the mass of abrasive media that develop the kinetic
energy that is released onto the workpiece surface. It follows that based on the abrasive media type,
the abrasion potential of different technologies is determined by the impact speeds/force between the
abrasive media and the workpiece. The energy of such interactions is converted to strain energy once
impacted on the workpiece surface.
Furthermore, the strain energy release must be high enough in order to achieve a critical value
equal to the resistance of crack growth, thereby leading to surface plastic deformations by micro-cutting.
This higher energy should be accurately controlled to avoid problems such as the over-finishing of
the workpiece surface. Over-finishing may cause a failure of the workpiece due to the massive
surface erosion.
The reduced residual material after machining, low energy requirements, and quick cycle times
make the fixed part processes of drag and stream industrially sustainable and promising in different
manufacturing areas where high quality and performance are crucial requirements.
In this study, an experimental investigation has been undertaken with statistical analysis to
determine the effect of varying DSS, DHS and media in the drag finishing process on the surface
roughness values of AM components, using the Taguchi method. Taguchi analysis is responsible for
investigating the rank and the level of factors that have a significant effect on the response. Taguchi
statistical analysis of the drag finishing process suggested that the control factors that minimise surface
roughness were ceramic media, DHS (level 2), 80 rpm and DSS (level 2), 50 rpm.
The study has identified the need to further characterise selected high-energy mass finishing
processes for more realistic, and hence potentially more challenging, AM parts. The study has
highlighted the potential for the future development of a two-stage finishing system.

Author Contributions: The research was conceived and completed jointly by Mikdam Jamal and
Michael N. Morgan. The Manufacturing Technoogy Centre and partners, and Liverpool John Moores University
technical staff assisted in the completion of experiments and surface measurements.
Conflicts of Interest: The authors declare no conflicts of interest.

85
Inventions 2017, 2, 36

References
1. Gillespie, L.K. Mass Finishing Handbook; Industrial Press: New York, NY, USA, 2006; ISBN 9780831132576.
2. Mediratta, R.; Ahluwalia, K.; Yeo, S.H. State-of-the-Art on Vibratory Finishing in the Aviation Industry:
An Industrial and Academic Perspective. 2015. Available online: link.springer.com/article/10.1007/s00170-
015-7942-0 (accessed on 1 July 2017).
3. Salvatorea, F.; Grangea, F.; Kaminskia, R.; Claudina, C.; Kermoucheb, G.; Recha, J.; Texierc, A. Experimental
and Numerical Study of Media Action during Tribofinishing in the Case of SLM Titanium Parts.
2017. Available online: www.sciencedirect.com/science/article/pii/S2212827117304341 (accessed on
1 August 2017).
4. Walton, K.; Blunt, L.; Fleming, L. The Topographic Development and Areal Parametric Characterization of
a Stratified Surface Polished by Mass Finishing. 2015. Available online: iopscience.iop.org/article/10.1088/
2051-672X/3/3/035003 (accessed on 1 August 2017).
5. Zaki, N. Mass finishing considerations for optimum productivity. Metal Finish. 1992, 90, 50–54.
6. Brust, P.C. Surface Improvement by Vibratory Cascade Finishing Process; Transactions-North American
Manufacturing Research Institution of SME: Lincoln, NE, USA, 1997; pp. 93–98.
7. Davidson, D.A. Developments in mass finishing technology. Metal Finish. 2003, 101, 49–56. [CrossRef]
8. Domblesky, J.; Evans, R.; Cariapa, V. Material removal model for vibratory finishing. Int. J. Prod. Res.
2004, 42, 1029–1041. [CrossRef]
9. Davidson, D.A. Surface finishing reaches new heights: Mass media finishing techniques can improve aircraft
part performance and service life. Metal Finish. 2005, 103, 25–28. [CrossRef]
10. Cariapa, V.; Park, H.; Kim, J.; Cheng, C.; Evaristo, A. Development of a metal removal model using spherical
ceramic media in a centrifugal disc mass finishing machine. Int. J. Adv. Manuf. Technol. 2008, 39, 92–106.
[CrossRef]
11. Telljohann, D. Stream Finishing Process and Application. 2014. Available online: http://www.otec.de/
(accessed on 1 August 2017).
12. Moore, M.A.; Douthwaite, R.M. Plastic deformation below worn surfaces. Metall. Trans. A 1976, 7, 1833–1839.
[CrossRef]
13. Domblesky, J.; Cariapa, V.; Evans, R. Investigation of vibratory bowl finishing. Int. J. Prod. Res. 2003, 41,
3943–3953. [CrossRef]
14. De Pellegrin, D.V.; Stachowiak, G.W. Assessing the role of particle shape and scale in abrasion using
‘sharpness analysis’: Part II. Technique evaluation. Wear 2002, 253, 1026–1034. [CrossRef]
15. Phadke, M.S.; Kackar, R.N.; Speeney, D.V.; Grieco, M.J. Off-line quality control in integrated circuit fabrication
using experimental design. Bell Syst. Tech. J. 1983, 62, 1273–1309. [CrossRef]
16. Pignatiello, J.J., Jr. An overview of the strategy and tactics of Taguchi. IIE Trans. 1988, 20, 247–254. [CrossRef]
17. Pease, G. Taguchi Methods: A Hands-On Approach to Quality Engineering; Addison Wesley: New York, NY,
USA, 1993.
18. Jamal, M.; Morgan, M.N.; Peavoy, D. A digital process optimisation, process design and process informatics
system for high energy abrasive mass finishing. Int. J. Adv. Manuf. Technol. 2017, 92, 303–319. [CrossRef]

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

86
inventions
Article
Direct Photonic Fusion of Vitrified Bonding Materials
Mark J. Jackson 1,2, * and Martin J. Toward 1
1 Department of Engineering, University of Liverpool, Liverpool L69 3BX, UK; [email protected]
2 Bonded Abrasive Group, Cambridge, Massachusetts, MA 01239, USA
* Correspondence: [email protected]; Tel.: +1-785-342-4784

Received: 17 July 2017; Accepted: 16 August 2017; Published: 18 August 2017

Abstract: The purpose of this study is to show the effects of the direct fusion of raw materials used for
vitrified grinding wheels by photonic interactions. The paper describes the construction of a sintering
apparatus that employs a pulsed neodymium: yttrium aluminum garnet (Nd:YAG) laser to fuse a
combination of raw materials such as ball clay, feldspar, and borax to form a partially-crystalline
glass material. The experimental results show that lasers can replace traditional methods of glass
frit formation by fusing raw materials used in the manufacture of glass bonds for vitrified grinding
wheels. X-ray diffraction data shows that a glass with short range order has formed using the
new method. The work described herein provides a new avenue for glass frit formation applied to
grinding wheel manufacture.

Keywords: grinding wheels; materials; bonding; glass; glass–ceramic; laser; fusion

1. Introduction
When studying the structure of vitrified bonds (i.e., raw materials that vitrify and form a
mechanical and chemical bond with abrasive grains), it is obvious that differences in bonding
characteristics are common in grinding wheels made with the same batch formulation but made
from a variety of commercially available raw materials. This is not surprising considering that clays
and feldspar minerals vary in composition and have different concentrations of trace elements.
Oxides present in clays and feldspars can affect crystallization behavior to a large degree [1].
The subsequent melting and devitrification behavior of these raw materials can greatly affect the
products of fusion. One way to prepare grinding wheels is to use pre-fitted raw materials where
oxide compositions can be tailored to suit a particular grinding operation [2]. The current method
of pre-fritting is to melt clays and feldspars at high temperature in a specially prepared platinum
crucible so that the molten mass can be quickly quenched from high temperature to prepare the melt
for fritting [3–6]. This is done by pouring the molten glass into a suitable coolant to control the rate
of glass formation prior to milling and crushing to form the powdered glass frit [7–10]. The main
aim of this work is to increase the rate at which the bonding materials are fused to form a glass or
glass–ceramic bond using a directed photon beam, which not only prepares the fused mass for fritting,
but can be further developed to form a vitrified network in a grinding wheel without the need for
fritting, thus reducing the time it takes to produce a commercially available grinding wheel.

2. Experimental Apparatus and Materials


A specially designed photonic sintering unit was constructed using a Rofin Sinar 90 W flash-lamp
pumped, Q-switched yttrium aluminum garnet (Nd:YAG) laser (molecular vibrating) with a frequency
range of 0–60 kHz (photon wavelength, λ = 1.06 μm (infra-red), photon frequency, ν = 2.8 × 1014 Hz,
quantum efficiency ~40%). The laser consisted of a standard cavity design with the active medium
being Nd3+ ions in a YAG crystal rod mounted to one of the foci of an elliptical chamber made from
gold-coated aluminum. At the opposing end was a krypton lamp whose emission spectrum suits

Inventions 2017, 2, 19; doi:10.3390/inventions2030019 87 www.mdpi.com/journal/inventions


Inventions 2017, 2, 19

the Nd3+ activity profile. Mode control was activated by a Q-switch that creates controlled photonic
oscillations in such a way that a mechanical chopper, optoelectronic switch, or an acoustic-optic switch
can be used to control the mode of the light beam [11]. For this particular laser, the acoustic-optic
switch propagates sound waves through a crystal lattice generated by a piezo-electric transducer.
A radio frequency oscillator generates sound waves in the 24–27 MHz range that produces waves
with an alternating refractive index, causing it to act like an optical grating and stopping the action of
the laser.
Approximately 60–100 W of power is required to deflect a 60 W beam, so the Q-switch has to
be cooled during laser operation. The speed of the optical shutters allows energy to build up in the
cavity when the laser action is halted, thus creating very high peak power to be generated. For a
20 W Nd-YAG laser, Q-switch action may produce 1 mJ/pulse (pulse widths ~6 ns), resulting in the
generation of 100 kW peak power. The pulse energy and pulse widths used in the current work were
in the range of 0.2–3 mJ and 80–250 ns, respectively. For pulse frequencies between 15 and 60 kHz,
the output power ranged from 2–12 W (lamp currents ~8–20 A (Figure 1)).

Figure 1. Laser power as a function of pulse frequency and lamp current for the Rofin Sinar 90 W
flash-lamp, pumped, Q-switched yttrium aluminum garnet (Nd:YAG) laser.

Scan line overlap was ~50% of line width, and the peak power density was 1.4 GW/cm2 .
Line scanning was achieved using a scanning head galvanometer containing two thermally regulated
galvanometers giving a scanning speed range of 1–500 mm/s over an 80 mm × 80 mm scanning
area. The focal length of the imaging lens was 112 mm, giving a minimum spot size of 50 μm.
The experimental apparatus is shown in Figure 2a.

(a) (b)

Figure 2. (a) Photonic sintering apparatus showing Q-switched Nd-YAG laser plus lens and
scanning head galvanometer mounted directly above the build chamber; and (b) Schematic layout of
laser–material interaction design-of-experiments showing magnitude of lamp current versus beam
pulse frequency. Arrow indicates the first, or initial, location of interaction experiments.

88
Inventions 2017, 2, 19

The build chamber consists of a build cylinder and powder delivery cylinder, each of 100 mm
diameter powered by two linear stepping motors with a minimum step size of 250 nm and maximum
stepping rate of 380 steps/s. A thin layer of powder (10 mm thickness) was mounted on the
build platform so that laser–material interactions could be characterized during the experimental
procedure. The parametric variables investigated were frequency (0–60 Hz), scanning velocity
(~50–500 mm/s), lamp current (8–20 A), focal length (104–112 mm), and spot size (100−300 μm).
The system configuration enabled a matrix of 100 parametric values to be examined. The schematic
layout of the laser–material interaction experiments is shown in Figure 2b.
The materials used in the experiments are those typically used in vitrified grinding wheels and
include Devon ball clay (BC), Finnish flotation feldspar (FFF), borax frit (BX), and mixtures of all
three raw materials. The preliminary results showed that no interactions occurred between the pulsed
photonic beam and the ball clay up to 17.3 A, then vaporization occurred. Rapid heating of a powder
results in a shock wave traveling towards the laser beam which vaporizes as it leaves the surface at
velocities >105 m/s. However, on inspection of the ball clay, there was no evidence of powder blasting
or cratering on the clay surface. This situation is tolerable because ball clay is used in grinding wheel
manufacture to support abrasive particles together as a matrix while the fusible component of the
bond melts and fills the interstices between clay particle and abrasive grain.

3. Experimental Results
The experimental work focused on the effects of changing process variables such as laser spot
size, focal length, scanning velocity, lamp current, and optical frequency on the physical appearance
and crystallographic structure of the fused bonding materials. The bonding materials processed were
ball clay (BC), feldspar (FFF), borax frit (BX), and a mixture composed of 60 wt. % frit (BX), 20 wt. %
ball clay (BC) and 20 wt. % feldspar (FFF).
Initially, X-ray diffraction of the raw materials was performed to characterize each bond ingredient
using a Phillips 1710 X-ray generator with a 40 kV tube voltage and a 30 mA current. Monochromatic
Cu kα radiation, λ = 1.54060 Å (wavelength intensity ratio ~0.5, divergence slit angle ~0.5◦ , receiving
slit angle ~0.2◦ ), was employed in the analysis. A scanning speed of 2◦ per minute for diffraction
angles of 2θ between 15◦ and 60◦ was recorded by computer (start angle ~5◦ (2θ), end angle ~5◦ (2θ),
step size ~0.04◦ (2θ)). The spectrum of each raw material was analyzed and compared with known
spectra. Powder specimens were prepared by crushing in a mortar and pestle in preparation for
quantitative X-ray diffraction. To eliminate the requirement of knowing mass absorption coefficients
of ceramic samples for quantitative X-ray diffraction, Alexander and Klug [12] introduced the use of
an internal standard for powder X-ray analysis. The powder bed subjected to “X-rays” should give the
maximum diffracted intensity. The spectrum for ball clay (BC) is shown in Figure 3a. The spectrum
shows kaolinite (Al2 Si2 O5 (OH)4 ) at 2θ angles (plane) [d-value (α1 )] of 12.285◦ (001) [7.199 Å], 19.79◦
(020) [4.4826 Å], 24.815◦ (002) [3.5851 Å], 34.92◦ (-201) [2.5673 Å], 35.925◦ (200) [2.4978 Å], 37.65◦
(003) [2.3872 Å], 50.955◦ (004) [1.7907 Å], 59.855◦ (114) [1.544 Å], 63.895◦ (223) [1.4558 Å], 68.255◦
(062) [1.3730 Å], and 72.14◦ (204) [1.3083 Å]. The next major component of ball clay is quartz (SiO2 ),
and shows 2θ angles (plane) [d-value (α1 )] of 26.585◦ (101) [3.3503 Å], 40.25◦ (111) [2.2388 Å], 54.81◦
(202) [1.6736 Å], and 67.655◦ (212) [1.3837 Å]. Muscovite mica ((K, Na)Al2 (SiAl)4 O10 ) also appears in
the spectrum, but is of low measurable order.

89
Inventions 2017, 2, 19

(a) (b) (c)

Figure 3. Ball clay (BC) material sample: (a) X-ray diffraction pattern for 2θ angles showing
arbitrary counts (counts) versus 2θ angles (2T) in degrees; (b) Scanning electron micrograph at
500× magnification; and (c) Scanning electron micrograph at 1250× magnification.

The spectrum for Finnish Flotation Feldspar (FFF) is shown in Figure 4a. The major components
detected include quartz (SiO2 ) showing 2θ angles (plane) [d-value (α1 )] of 26.585◦ (101) [3.3503 Å],
40.25◦ (111) [2.2388 Å], 54.81◦ (202) [1.6736 Å] and 67.655◦ (212) [1.3837 Å] and albite (NaAlSi3 O8 )
showing reflections at 020, 003, 112, 040 and 004 planes with corresponding d-values of 6.42, 4.24, 3.756,
3.21 and 3.18 Å, respectively.

(a) (b) (c)

Figure 4. Finnish flotation feldspar (FFF) material sample: (a) X-ray diffraction pattern for 2θ angles
showing arbitrary counts (counts) versus 2θ angles (2T) in degrees; (b) Scanning electron micrograph
at 500× magnification; and (c) Scanning electron micrograph at 1250× magnification.

The spectrum for Borax frit (BX) is shown in Figure 5a. The major components include anhydrous
borax (Na2 B4 O7 ), which is approximately 31 wt. % Na2 O and 69 wt. % B2 O3 , and quartz (SiO2 )
showing 2θ angles (plane) [d-value (α1 )] of 26.585◦ (101) [3.3503 Å], 40.25◦ (111) [2.2388 Å], 54.81◦ (202)
[1.6736 Å], and 67.655◦ (212) [1.3837 Å].

(a) (b) (c)

Figure 5. Borax frit (BX) material sample: (a) X-ray diffraction pattern for 2θ angles showing
arbitrary counts (counts) versus 2θ angles (2T) in degrees; (b) Scanning electron micrograph at
500× magnification; and (c) Scanning electron micrograph at 1250× magnification.

The bonding materials were then handled as powders loaded manually in the build chamber to
ensure direct photonic interactions between laser beam and powder bed as a single layer presented on

90
Inventions 2017, 2, 19

a planar surface (Figure 6). The raw materials were exposed to the photonic beam, which projected
itself as a spot of various sizes (100, 200, and 300 μm) scanned across the powder bed at various
speeds—typically 50, 250, and 500 mm/s.

Figure 6. The build chamber showing a thin layer of powder presented as a single layer on a planar
surface prior to direct photonic interaction.

The following observations were made for each material after a thin layer was exposed to the
laser beam at various spot sizes and laser scan velocities:
(i) Ball clay (BC)—for a 100 μm spot size and 50 mm/s scan velocity, there was no interaction
between laser and ball clay up to 9.8 A lamp current, beyond 9.8 A and at a frequency of 26 kHz,
burning of the ball clay occurred. When the scan velocity was increased to 250 mm/s, vaporization
began at 18.6 A lamp current between 0 and 60 kHz. Beyond 500 mm/s scan velocity, no interaction
was observed between the photonic beam and the ball clay. For 200 μm spot size, slight burning of the
ball clay powder occurred at 16.9 A lamp current between 0 and 60 kHz frequency; at larger spot size
(~300 μm), no interactions occurred at higher scan velocities, although burning occurred at 50 mm/s
scan speed at 16.9 A lamp current between the 0 and 60 kHz frequency range. The topography of the
BC powder bed is shown in Figure 3b,c at 500× and 1250× magnifications, respectively.
(ii) Finnish floatation feldspar (FFF)—at 100 μm spot size and 50 mm/s scan velocity, discoloration
of the powder occurred up to 10.6 A lamp current between 0 and 60 kHz frequency, after which
vaporization occurred. Above 33 kHz frequency, the powder appeared to crystallize. This observation
occurred at higher scan velocity until a 20 A lamp current produced complete crystallinity. With large
spot sizes, crystalline phases were created at lamp currents >18.6 A. For a 300 μm spot size, FFF
crystallinity occurred above 17.8 A lamp current between 0–60 kHz frequency range. For higher scan
velocities (>250 mm/s), FFF became white in color. The topography of the FFF powder bed is shown
in Figure 4b,c at 500× and 1250× magnifications, respectively.
(iii) Borax frit (BX)—no interactions were observed up to 11.9 A lamp current through the
frequency range of 0–60 kHz. Beyond 11.9 A lamp current, slight fusion of the frit occurred at
low scan velocities. The topography of the BX powder bed is shown in Figure 5b,c at 500× and
1250× magnifications, respectively.
(iv) Sintered bond mix—at 100 μm spot size and 50 mm/s scan velocity, no interactions were
observed up to 9.8 A lamp current in the 0–60 kHz frequency range. Between 9.8 and 11.9 A lamp
current, vaporization occurred with slight crystallinity observed in the bond mix. At lamp currents
beyond 13.8 A, burning of the mix occurred. At higher scan velocities, partial fusion or no interactions
occurred. Similar observations were made with large spot sizes, the optimum conditions being lamp
currents between 9.8 and 11.9 A in the 0–60 kHz frequency range to produce vaporization of the
mix accompanied by slight crystallinity. The topography of the bond mix powder bed is shown in
Figure 7b,c at 500× and 1250× magnifications, respectively. The expected equilibrium phases from
the directly fused mixture (20 wt. % BC, 20 wt. % FFF, and 60 wt. % BX) are quartz (unreacted and
partially dissolved), mullite, cristobalite, and glass. However, from the samples tested, the compounds
quartz, mullite, and glass were successfully detected, indicating non-equilibrium formation conditions

91
Inventions 2017, 2, 19

(Figure 7a). A calibration curve was constructed using a suitable internal standard (CaF2 ), a diluent,
and a synthetic form of the phase(s) to be measured. Synthetic mullite with purity >99.8% and
powdered quartz with purity >99.84% SiO2 were used throughout the experiments. The method used
for quantitative analysis of ceramic powders was developed by Khandelwal and Cook [13] and refined
by Monshi and Messer [14].
The internal standard provides an intense (111) reflection (d = 1.354 Å) lying between the (100)
reflection for quartz (d = 4.257 Å) and the (200) reflection for mullite (d = 3.773 Å). Using copper kα
radiation (λ = 1.5405 Å), the corresponding values of diffraction angle 2θ are: (100) quartz = 20.82◦ ;
(111) calcium fluoride = 28.3◦ ; and (200) mullite = 32.26◦ . The calibration curve generated by varying
proportions of calcium fluoride, synthetic quartz, and mullite is shown in Reference [8]. Mass fractions
of the crystalline phases in the mixture can be interpreted from the calibration lines by measuring the
intensity ratio of the phase(s) to the internal standard. The diffraction peaks of interest for quantitative
analysis lie between 15◦ and 40◦ of the diffraction angle 2θ [8]. The reflections are the (111) planes of
calcium fluoride, (200) planes of mullite, and the (100) planes of quartz. In order to calculate the mass
fractions of quartz and mullite in the mixture, the height of the chosen diffraction peak and its width
at half-height were measured from the diffraction spectrum. The product of these two measures were
then compared with that of the internal standard, and the resultant intensity ratio was used to find the
exact mass fraction of the phase(s) measured in the glass that was subjected to X-ray diffraction [8].

(a) (b) (c)

Figure 7. Sintered bond mix (20 wt. % BC, 20 wt. % FFF, and 60 wt. % BX) material sample:
(a) X-ray diffraction pattern for 2θ angles showing arbitrary counts (counts) versus 2θ angles (2T) in
degrees; (b) Scanning electron micrograph at 500× magnification; and (c) scanning electron image at
1250× magnification.

The X-ray spectrum for bond mixture is shown in Figure 7a. The major components detected
include quartz (SiO2 ) showing 2θ angles (plane) [d-value (α1 )] of 26.585◦ (101) [3.3503 Å], 40.25◦ (111)
[2.2388 Å], 54.81◦ (202) [1.6736 Å], and 67.655◦ (212) [1.3837 Å], mullite (Al6 Si2 O13 ) showing reflections
at (110), (200), (001), (111), and (201) planes with corresponding d-values of 5.39, 3.774, 2.886, 2.542,
and 2.292 Å, respectively, and a non-crystalline glass phase. The images shown in Figure 7b,c show
that the surrounding bond mix has melted, while large clay relicts have remained intact. The bubbles
shown are decomposed gases held in the clays that form and escape from their matrix during melting.
These gases are typically CO2 and SO2 , and are seen in molten glass frits during standard grinding
wheel manufacture. The release of gases contained in clays and various pore forming agents occurs at
high temperatures (>650 ◦ C). The images shown in Figure 7 and the X-ray spectra of the bond mix
subjected to laser beam interactions are not dissimilar to that of a glass frit made using the traditional
way of melting where bonding materials are placed in a platinum crucible, heated to the melting
temperature, then poured in a cooling furnace for the preparation of fritting prior to use in vitrified
grinding wheels.

92
Inventions 2017, 2, 19

4. Discussion
The fusing of the raw materials was apparent at low scanning speeds and spot sizes. The powder
formed a brittle structure that was full of porosity. The depth of fusion was ~0.5 mm for the bond
mix. However, ball clay (BC) scanned separately burned at low scan speeds and spot sizes. At higher
levels, no interaction between the photonic beam and the clay occurred. This is advantageous for
the preparation of frits for grinding wheels because the clay particles remain intact as the lower
melting point material flows around it. It is noted that the melting point of ball clay (BC) is ~1400 ◦ C.
Feldspar (FFF) particles behaved a little differently. Slight melting occurred at the edges of the feldspar
particles at low scan speeds and spot sizes (FFF melting point is ~1200 ◦ C). Although some melting is
encouraged for feldspar, complete melting is not required in the bond mix because a large size grinding
wheel may collapse under its own weight when fired due to the melting of too many bond components
at the same temperature. The most surprising of all the raw materials was the behavior of borax frit
(BX) during photonic interactions. The melting point of borax frit is ~742 ◦ C, so one should have
anticipated melting during the photonic interactions. However, direct interaction with BX frit tends to
indicate that the photon beam is reflected by the thin powder layer of the BX frit. When mixed with
the BC and FFF, BX frit appears to melt and form a highly non-crystalline structure that is shown in the
X-ray spectrum shown in Figure 7a. This tends to indicate that the heat absorbed by BC and FFF by
exposure to the photonic beam may be responsible for heating and causing BX to melt when presented
as a mixture to the laser beam. This presents a very interesting situation in terms of providing raw
materials for vitrified grinding wheels, because materials that reflect optical energy may be blended
with materials that absorb energy in order to provide homogeneous melting of the bond mixture that
bonds to abrasive grains. This implies that abrasive grains may not require heating during grinding
wheel manufacture that uses lasers, thus preserving the initial strength and mechanical characteristics
of abrasive grains that are normally impaired due to long periods of traditional thermal treatment
in kilns.
Future research may therefore lead to the development of grinding wheels whose abrasive
grains maintain their original properties due to the absence of thermal degradation caused by the
use of focused photonic energy to selectively heat parts of the grinding wheel during manufacture.
The avoidance of using clays altogether may be possible for small grinding wheels, because clays may
be replaced by oxides whose composition may be tailored to suit the particular grinding application.
However, the fusibility of oxides must be examined to ensure that photonic interactions can truly
achieve the production of glass frits without the use of clays. It is expected that experiments described
in this paper will lead to the development of new inventions associated with the production of glass
bonding systems and new innovations in the way that direct photonic fusion can be used in the
manufacture of vitrified grinding wheels. New inventions will be focused on producing vitrified
grinding wheels that vitrify in a shorter amount of time using much less energy than current methods.
The current work has shown that not only is direct photonic fusion of vitrified bond materials possible,
but when applied to grinding wheel manufacture, the process may preserve the fracture characteristics
of virgin abrasive grains without the deleterious effects of thermal degradation caused by traditional
methods of grinding wheel manufacture.

Acknowledgments: The authors thank Paul Dando of Saint-Gobain Abrasives for supplying the raw materials
and John Curran for training Martin Toward to use the laser sintering apparatus and its construction. The authors
also thank Professor Bernard Hon for allowing the authors to conduct the work at the Product Innovation
Development Center, University of Liverpool. The authors thank the Engineering and Physical Science Research
Council for funding this work under grant number GR/N06731/01.
Author Contributions: Mark J. Jackson and Martin J. Toward conceived and designed the experiments;
Martin J. Toward performed the experiments; Mark J. Jackson and Martin J. Toward analyzed the data; and
Mark J. Jackson wrote the paper.
Conflicts of Interest: The authors declare no conflict of interest.

93
Inventions 2017, 2, 19

References
1. Jackson, M.J.; Hitchiner, M.P. Vitrified Bonding Systems and Heat Treatment. In High Performance Grinding
Advanced Cutting Tool; Springer Nature: New York, NY, USA, 2013; pp. 45–94.
2. Graf, W. Overview of Abrasives. In Handbook of Creep-Feed and Surface Grinding; Winterthur Schlieftechnik
AG: Winterthur, Switzerland, 2010; p. 21.
3. Jackson, M.J.; Mills, B. Interfacial bonding between corundum and glass. J. Mater. Sci. Lett. 2000, 19, 915–917.
[CrossRef]
4. Jackson, M.J. Studies on refractory bonding systems used in vitrified silicon carbide grinding wheels.
Proc. Inst. Mech. Eng. 2000, 214, 211–221. [CrossRef]
5. Jackson, M.J.; Mills, B. Vitrification heat treatment and the dissolution of quartz in grinding wheel bonding
systems. J. Inst. Mater. Miner. Min. 2001, 100, 1–8. [CrossRef]
6. Jackson, M.J.; Wakefield, R.T.; Jones, S.A.; Mills, B.; Rowe, W.B. Materials selection applied to vitrified
corundum grinding wheels. J. Inst. Mater. Miner. Min. 2001, 100, 229–236.
7. Jackson, M.J.; Mills, B. Microscale wear of vitrified abrasive materials. J. Mater. Sci. 2004, 39, 2131–2143.
[CrossRef]
8. Jackson, M.J. Tribological design of grinding wheels using X-ray diffraction techniques. Proc. Inst. Mech. Eng.
2006, 220, 1–17. [CrossRef]
9. Jackson, M.J. Sintering and vitrification heat treatment of cBN grinding wheels. J. Mater. Proc. Technol. 2007,
191, 232–234. [CrossRef]
10. Jackson, M.J. An historical review of wear mechanisms and the structure of vitrified grinding wheels.
Int. J. Nanomanuf. 2009, 3, 368–397. [CrossRef]
11. Steen, W.M. Chapter 1—Background and General Applications. In Laser Materials Processing, 2nd ed.;
Springer: Berlin, Germany, 1998; pp. 11–57.
12. Alexander, L.; Klug, H.P. Basic aspects of X-ray absorption. Anal. Chem. 1948, 20, 886–889. [CrossRef]
13. Khandelwal, S.K.; Cook, R.L. Effect of alumina additions on crystalline constituents and fired properties of
electrical porcelain. Am. Ceram. Soc. Bull. 1970, 49, 522–526.
14. Monshi, A.; Messer, P.F. Ratio of slopes method for quantitative X-ray diffraction analysis. J. Mater. Sci. 1991,
26, 3623–3627. [CrossRef]

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

94
inventions
Article
Grinding Fluid Jet Characteristics and Their Effect
on a Gear Profile Grinding Process
Philip Geilert 1, *, Carsten Heinzel 1,2 and André Wagner 1
1 Stiftung Institut für Werkstofftechnik (IWT), Badgasteiner Strasse 3, D-28359 Bremen, Germany;
[email protected] (C.H.); [email protected] (A.W.)
2 MAPEX Center for Materials and Processes, University of Bremen, Bibliothekstrasse 1,
D-28359 Bremen, Germany
* Correspondence: [email protected]; Tel.: +49-421-218-51184

Received: 26 September 2017; Accepted: 20 October 2017; Published: 25 October 2017

Abstract: Profile gear grinding is characterized by a high level of achievable process performance and
workpiece quality. However, the wide contact length between the workpiece and the grinding wheel
is disadvantageous for the fluid supply to the contact zone and leads to the risk of locally burning
the workpiece surface. For the reduction of both the thermal load and the risk of thermo-mechanical
damage, the usage of a grinding fluid needs to be investigated and optimized. For this purpose,
different kinds of grinding fluid nozzles were tested, which provide different grinding fluid jet
characteristics. Through a specific design of the nozzles, it is possible to control the fluid flow inside
the nozzle. It was found that this internal fluid flow directly influences the breakup of the coolant
fluid jet. There are three groups of jet breakup (“droplet”, “wave & droplet”, and “atomization”).
The first experimental results show that the influence of the jet breakup on the process performance
is significant. The “wave & droplet” jet breakup can achieve a high process performance, in contrast
to the “atomization” jet breakup. It can therefore be assumed that the wetting of the grinding wheel
by the grinding fluid jet is significantly influenced by the jet breakup.

Keywords: profile gear grinding; grinding fluid; grinding fluid nozzle; jet breakup

1. Introduction and the State of the Art


In grinding processes in general, a high risk of thermo-mechanical damage exists due to the
kinematics of the abrasive grain in the contact zone as well as the large contact area between the
grinding wheel and the workpiece. This can lead to residual tensile stresses and changes in hardness,
which can have a negative effect on the lifetime of the component [1,2]. Especially in profile gear
grinding processes, the large contact area between grinding wheel and workpiece makes it difficult
to supply the grinding fluid to the contact zone. To avoid thermo-mechanical damage, in most cases,
the supply of a grinding fluid is indispensable and typical in industrial applications. Due to a lack of
experience in appropriately supplying grinding fluid in practice, high grinding fluid flow rates are
normally chosen to achieve high process reliability.
Experience has shown that the productivity of the process, workpiece quality, and tool wear can
be influenced not only by the chemical composition of the grinding fluid, but also by the grinding
fluid supply. In addition, large grinding fluid reservoirs are needed when high grinding fluid flow
rates are used, because of the necessary cooling and settling phases of the grinding fluid. Furthermore,
high grinding fluid flow rates and grinding fluid pressures result in a high level of aerosol pollution,
necessitating an exhaust system to ensure the health of the machine operator [3]. Therefore, the aim
should be to reduce the grinding fluid volume without affecting the workpiece quality and the
material removal rate. Thus, a high performance grinding fluid supply becomes particularly important.

Inventions 2017, 2, 27; doi:10.3390/inventions2040027 95 www.mdpi.com/journal/inventions


Inventions 2017, 2, 27

Consequently, a grinding fluid supply should be developed on the basis of “as much as necessary—as
little as possible” [4–7].
A nozzle contour that is adapted to the profile gear grinding wheel contour can considerably
increase the efficiency of the grinding fluid supply. In addition, the jet velocity and the coherence of
the jet are important in grinding fluid nozzle design. To break through the air belt surrounding the
grinding wheel, an adapted jet velocity is necessary. When the circumferential speed of the grinding
wheel and the velocity of the grinding fluid jet are nearly identical, the wetting of the grinding wheel
increases. In contrast to widening jets, a coherent jet carries less air into the grinding contact zone,
and a larger amount of grinding fluid flows through the grinding contact zone [8]. As a result of these
adjustments, the grinding fluid flow rate can be significantly decreased without any negative effect on
the workpiece surface.
In addition, the nozzle also influences the jet characteristics and the jet breakup in particular.
While studies have been conducted in the research fields of “atomization technology” and “combustion
engines”, focusing on the nozzles and their jet breakup behavior, no research work is known that deals
with grinding fluid nozzles and their influence on both jet breakup and the grinding processes.

2. Research Approach and Objective


The aim of the presented work is to study the influence of the jet breakup on the grinding process.
Therefore, the relationship between the nozzle designs and the jet breakup characteristics is analysed.
Dimensionless numbers like the Ohnesorge number and the Reynolds number are used to classify the
nozzles with regard to their jet characteristics. In a second step, it is attempted to establish a relationship
between these theoretical investigations and the jet breakup characteristics, as well as the influence on
the grinding process. Therefore, a gear profile grinding process is used and the nozzles are compared
with regard to their influence on a thermo-mechanical damage of the workpiece. This should provide
a comprehensive understanding of the relationship between the grinding fluid nozzle design, the jet
breakup, and the grinding process.
The gained scientific knowledge should allow grinding fluid nozzles to be designed for an optimal
wetting of the contact zone in profile grinding. As a result, the performance and energy efficiency of
gear grinding processes can be expected to increase.

3. Materials and Methods


In the following, the theoretical backgrounds for designing grinding fluid nozzles, as well as
the different nozzles used for the experimental investigations are described. This is followed by
a description of the methods used to characterize a fluid jet. Furthermore, the machine and workpiece
used for the grinding experiments are specified, and the micromagnetic test for thermo-mechanical
damage is also introduced.

3.1. Nozzle Designs


A jet, which can normally be described by a cylindrical volume, decays because of the instability
that is driven by capillary forces. This is known, amongst other things, as the Plateau-Rayleigh-
Instability. This effect explains how a jet decays into a particle chain as it progresses and breaks apart
until it takes the form with the smallest surface energy, the droplet [9].
In this context, three dimensionless numbers are important to characterize a fluid jet. In particular,
the jet breakup can be characterized in advance in the Ohnesorge vs. Reynolds number diagram.
The jet breakup can be divided into three fields (“droplet” (a), “wave & droplet” (b), and “atomization”
(c)) (Figure 1) [9,10]. Due to the high jet velocity, the fields (b)] and (c) are of importance for grinding
fluid nozzles.
The Reynolds number (Re) is a dimensionless number that is used to predict the transition from
laminar to turbulent flow.

96
Inventions 2017, 2, 27

ρ·v· D
Re = H
 η
ρ : oil density g/cm v : mean velocity[m/s]
3 (1)
D H : hydraulic diameter[mm] η : oil dynamilc viscosity [Pa·s]
The Ohnesorge number (Oh) relates the viscous forces to the inertial and surface tension forces:

η
Oh = √ = We
Re
D H ·ρ·σ (2)
 
σ : oil surface tension kg/s2

Thereby, it provides information about the characterization of the fluid atomization. The lower
the Ohnesorge number, the weaker the friction losses due to viscous forces. This means that most of
the inserted energy converts into surface tension energy and a droplet can be formed. The higher the
Ohnesorge number, the more dominant is the internal viscous dissipation and the droplet decays into
smaller particles [9,10].
The Ohnesorge number can be expressed as the square root of the Weber number divided by
the Reynolds number (Equation (2). The Weber number is described in terms of the inertia force, Fρ ,
divided by the surface tension, Fσ :
Fρ ρ · v2 · D H
We = = (3)
Fσ σ
By means of the Weber number, it can be evaluated how far the real shape of a droplet differs
from a spherical shape. Thereby, it is a measure for the deformation of a droplet, which increases with
an increasing Weber number. Therefore, the droplet will breakup into more and more tiny droplets
with an increasing Weber number [11].

Ohnesorge vs. Reynolds number diagram


10-1
101
(b) (c) {1}
(a) category
10-2 “atomization”
10-1 category {2}
“wave &
droplet”
{3}
{5} {4}
10-3 10-3
103 104 105

101 103 105 nozzle types

{1} needle nozzle


jet breakup characteristics {2} flat nozzle*
atomization {3} 3D printed A
wave & droplet {4} solid stream nozzle*
droplet {5} 3D printed B*

*designed and built by IWT

Figure 1. Ohnesorge vs. Reynolds number diagram and jet breakup characteristics (after [10]).

97
Inventions 2017, 2, 27

Based on this, grinding fluid nozzles {1–5} were chosen, which have an identical outlet cross-
section and therefore provide the same average jet velocities (vjet = 35 m/s) at the outlet for the same
flow rates (Qf = 100 l/min) (Figure 2). For comparison, the reference nozzle {r} was also investigated.
This nozzle represents the current state of technology in the industrial environment. This nozzle
differs from the nozzles {1–5} in the grinding fluid flowrate (Qf = 330 l/min) and in the jet velocity
(vjet = 12 m/s). Due to the low jet velocity, the jet does not break up before it reaches the grinding
contact zone.
To determine the positions for the nozzles {1–5} in the Ohnesorge vs. Reynolds number diagram
and the characteristics of the jet breakup, the dimensionless numbers are calculated. The oil used
has a surface tension of σ = 0.03 kg/s2 , a dynamic viscosity of η = 8.035 Pa·s, and a density of
ρ = 0.837 g/cm3 . Using the mean velocity and the hydraulic diameter in the nozzle, which are specific
to each nozzle, the Ohnesorge and the Reynolds numbers are calculated. The results are shown
in Figure 1.
For these settings, the nozzles can be classified into three categories. The needle nozzle {1},
the flat nozzle {2}, and the three-dimensional (3D) printed nozzle A {3}, which are close to the area
“atomization” in the diagram (category “atomization”). The solid stream nozzle {4} and the 3D printed
nozzle B {5} are in the area “wave & droplet”, close to the area “droplet” (category “wave & droplet”).
The reference nozzle {r} generates a coherent jet with no breakup for the considered conditions (category
“coherent jet”) at a rather small jet velocity far below cutting speed vc .

Figure 2. Different concepts of fluid supply nozzles.

3.2. Characterization of Fluid Dynamic Aspects of Grinding Fluid Jets


The grinding fluid jet breakup was characterized with the help of high speed photography.
A camera was used to take high resolution pictures (20.2 mega pixel) with an extremely short exposure
time. The pictures were illuminated from behind the grinding fluid jet, whereby a flash illuminated
a translucent glass plate with an exposure time of 1/60,000 s. The even illumination of the background
is transmitted through the grinding fluid jet. The pictures with the transmitted light through the
grinding fluid jet allow knowledge to be gained regarding the jet characteristics, namely the widening
of the jet, the distribution of the droplets, and the jet breakup (Figure 3).

Figure 3. Measurement setup for high-speed photography.

98
Inventions 2017, 2, 27

The impact pressure of the grinding fluid jet was measured using a high resolution piezoelectric
sensor (measuring tip diameter: 1.0 mm). The distribution of pressure was determined by moving
the sensor through the grinding fluid jet (Figure 4, bottom right). A flow-optimized cover (Figure 4,
top right) redirected the jet behind the measurement area, so that no deflected fluid could influence the
measurement. These measurements allow for a better understanding of the distribution and the size of
the droplets inside the jet. Thus it is possible to review the category of jet breakup for the nozzles.

Figure 4. Experimental setup, scanning program and repeatability for impact pressure measurement.

The sensor scanned the grinding fluid jet ten times on one line and measured the force at several
measuring points. The error bars indicate the min–max values (Figure 4, bottom left) and show
reproducible results. In addition to the distribution of pressure, the time sequences were analyzed and
it was possible to obtain knowledge about the regularity and the height of the impact pressure of the
droplet chain.
The energy efficiency of the nozzles is evaluated with the static pressure within the pipe directly
in front of the nozzle. The pressure is measured with a calibrated digital pressure transmitter. The used
silicon sensor has a resolution of 0.1 bar.

3.3. Grinding Process


In the profile grinding experiments, pre-machined case-hardened gears were machined on a gear
grinding machine (KAPP KX 500 FLEX), which enabled the machining of gears with a diameter of
up to 500 mm as well as a module from m = 0.5 to 10 mm with conventional and superabrasive
tools (Figure 5).
The workpieces are helical gears. These gears are characterized by 47 teeth, a normal module of
mn = 4.5 mm, a bevel angle of β = −16.55◦ , a pressure angle of α = 24◦ , and a width of b = 65 mm.
The material is AISI 5120 and the gears are case-hardened and blast-cleaned. The gears have a hardness
of 718 HV and a case hardness depth Chd of 1.13 mm.

99
Inventions 2017, 2, 27

Figure 5. Experimental setup for grinding tests.

For the micromagnetic test to characterize the thermo-mechanical impact on the ground tooth
flanks, the Stresstech measuring device Rollscan R 300 and the Stresstech NC unit GearScan 500 were
used (Figure 6). These instruments enable a fully automatic testing of the gear. Hereby, a sensor moves
by means of NC-controlled axes on previously programmed tracks along the tooth flank. Barkhausen
noise was analyzed at the center of the right tooth profile. This automatic measurement allows a high
reproducibility in comparison to manual tests.

Figure 6. GearScan 500 in use and measurement parameters.

4. Results and Discussion


As explained below, the fluid dynamic investigations provide a characterization of fluid jets and
their breakup. Furthermore, the grinding experiments show the impact of the jet characteristics on the
grinding process.

4.1. Characterization of Fluid Jets and Their Breakup


Through the use of the above-described high-speed photography, the fluid jets of the different
nozzles were characterized. A standardized measurement of the length of the coherent part of the jet

100
Inventions 2017, 2, 27

lcoherent before atomization makes it possible to compare the fluid jets. The results for these lengths,
which are the average lengths of nine measurements of each nozzle, are shown in Figure 7.

Figure 7. Characterization of fluid supply nozzles—coherent jet lengths.

The jets of the needle nozzle {1}, the flat nozzle {2}, and the 3D printed nozzle A {3} decay within
a short length. The three values for the coherent length are similar to each other, as well as the values
in the Ohnesorge vs. Reynolds number diagram are closely matched. The dark areas in the photos
show many tiny droplets, which adsorb the light of the illuminated background. This observation
points out that the three nozzles are comparable with regard to their position in the Ohnesorge vs.
Reynolds number diagram and generate an “atomization” jet breakup. This might be due to the high
turbulence directly at the nozzle outlets, which is caused by the high mean velocity inside the nozzles.
The solid stream nozzle {4} and the 3D printed nozzle B {5} have inner designs with optimized flow
characteristics, which significantly reduce the turbulence at the nozzle outlet. This can explain the
results for the longer coherent lengths of these two nozzles. For the evaluation of the coherent lengths,
only the respectively lower outlet was analysed, as defects in the 3D-printing of the upper outlet were

101
Inventions 2017, 2, 27

determined for the 3D printed nozzle B {5} (these defects result in a short coherent length for the
upper outlet). The results again reflect the relationship between the positions in the Re-Oh-diagram,
which are comparable for both nozzles {4} and {5}, and the coherent jet lengths for this category.
Apart the jet characteristics, the nozzles influence the measured pressure in the pipe directly in front
of the nozzle for a constant flow rate (Figure 7). For the needle nozzle {1} a pressure of pnozzle = 13.8 bar is
measured, whereby for 3D printed nozzle B {5} a pressure of pnozzle = 5.7 bar is measured. The nozzle {2}–{4}
are between these values. These pressure values correlate in inverse manner with the coherent lengths.
These different power drops in the nozzles generate a corresponding turbulence flow, which correlates
with the jet break up.
It can be summarized that the five nozzles {1–5} can be classified into the two categories
“atomization” and “wave & droplet”, and the results for the coherent jet lengths matches with this
observation. For the category “wave & droplet” greater values for the coherent length can be reached
than for the category “atomization”.
Besides the optical jet characteristics, the fluid jets of the different nozzles differ significantly
from each other with regard to the temporal course of the impact pressure. The impact pressure was
measured for a nozzle of each category. The results of the measurement for the reference nozzle {r}
(category “coherent jet”), as well as for the needle nozzle {1} (category “atomization”) and the solid
stream nozzle {4} (category “wave & droplet”) are shown in Figure 8.

ljet = 90 mm

atomization

*designed and built by IWT


vjet = 35 m/s
piezoelectric
probe
sensor
wave & droplet

vjet = 35 m/s
vjet = 35 m/s

category “coherent jet“ category “atomization“ category “wave & droplet“


reference nozzle needle nozzle solid stream nozzle*
impact pressure pimp [N/mm²]

impact pressure pimp [N/mm²]

impact pressure pimp [N/mm²]

min - Max
Min max = 3.65 mN
3,65 mN min - max = 11.46
, mN min - max
Min Max = 34.73
34,73 mN
mN
{r}
2


2ı = 1.18
1,18 mN
mN 5

1
{1} 2ı
2ı == 3.76
3,76 mN
mN
{4} 2ı == 8.93
8,93 mN

time t [s] time t [s] time t [s]

Figure 8. Characterization of the fluctuations in impact pressure for different jet breakups.

The reference nozzle {r} shows the lowest fluctuations in impact pressure. This is due to the low
jet velocity and the coherent jet. Despite the equal and constant fluid flow rate and the jet velocity of
the solid stream nozzle {4}, as well as the needle nozzle {1}, the solid stream nozzle {4} shows the most
significant fluctuations in impact pressure. This can be explained by the distribution of the droplets,
which is also shown in high-speed photos. The solid stream nozzle {4} generates a comparably long
coherent grinding fluid jet that decays in the transitional area into relatively large droplets, which
generates a greater impact pressure fluctuations than small ones. The fluid jet of the needle nozzle
{1} shows an “atomization” breakup of the jet close to the nozzle outlet. The resulting small droplets
are distributed evenly and therefore generate only minimal fluctuations, as well as a relatively low
impact pressure.

102
Inventions 2017, 2, 27

The results show that grinding fluid nozzles not only have an impact on the widening of the jet as
a function of jet velocity and flow rate, but also affect the jet breakup, the distribution of the droplets,
and therefore the fluctuations in impact pressure.

4.2. Grinding Technology


Different grinding fluid nozzles were applied that show different jet characteristics (“coherent
jet”, “wave & droplet” and “atomization”). In order to compare these types of nozzles, a grinding
process was developed that was tested three times with the reference nozzle {r} to determine the
reproducibility. For a practical process and a fast reaching of the process limit, the gears were machined
without pre-grinding in eight strokes and with a depth of cut ae of 50 μm in each case. The dressing
conditions were precisely adjusted to the requirements, so that an increased thermo-mechanical load
of the surface layer could be achieved after grinding only a few gear gaps. The results for this reference
process are shown in Figure 9.

Figure 9. Barkhausen noise and spindle power for the reference gear grinding process.

The courses of the spindle power in all three of the repetitions are similar to each other, and the
courses of the Barkhausen noise analysis are also closely matched. Hence, the repetition trials verify
the reproducibility of the process. In order to compare the grinding fluid supply conditions for the
six nozzles, the limit for the thermo-mechanical damage were compared using Barkhausen noise.
Therefore, the limit for the thermo-mechanical damage for the reference process was determined with
nital etching. When the nital etching indicates grinding burn, the Barkhausen noise value MP exceeds
60. For the reference process, a specific removed material volume V’W of 345 ± 15 mm3 /mm was
removed before this limit had been reached.
Based on the above-described grinding process, as well as the method to detect thermo-mechanical
damage, all discussed nozzles were analyzed. The values for the specific removed material volume V’W ,
before thermo-mechanical damage occurs, for each nozzle are shown in Figure 10.

103
Inventions 2017, 2, 27

Figure 10. Grinding tests—Influence of the jet breakup.

Using the reference nozzle {r}, a specific removed material volume of V’W = 345 ± 15 mm3 /mm was
reached. This result can be attributed to the much higher grinding fluid flowrate (Qf = 330 l/min) and
the coherent jet. For the nozzles of the category “atomization” and “wave & droplet”, a significant lower
grinding fluid flowrate of Qf = 100 l/min and a higher jet velocity of vjet = 35 m/s was adjusted which
corresponds to the cutting speed vc . With the needle nozzle {1}, the flat nozzle {2}, and the 3D printed
nozzle A {3} (category “atomization”), a specific removed material volume of V’W = 311–349 mm3 /mm
was achieved. The highest specific removed material volume of V’W = 415–436 mm3 /mm was reached
with the nozzles of the category “wave and droplet”. The results of the grinding tests correlate with
the coherent lengths of the jets and the calculated positions in the Ohnesorge vs. Reynolds number
diagram. A longer coherent jet leads to a higher specific removed material volume. This might be due
to a better wetting of the grinding wheel for a “wave and droplet” jet breakup. In conclusion, for the
optimized grinding fluid supply versus the reference fluid supply, the specific removed material volume
can significantly be increased (up to 26%//436 mm3 /mm {5} to 345 mm3 /mm {r}) with a simultaneous
decrease of the grinding fluid flow rate (reduction of 70%//100 l/min {5} to 330 l/min {r}). For a reduced
grinding fluid flow rate (100 l/min), an appropriate inner design of the nozzle alone can lead to an increase
in the process performance by up to 40% (436 mm3 /mm {5} to 311 mm3 /mm {1}).

5. Conclusions and Outlook


In the fluidic investigations conducted in this study, the grinding fluid jet and the jet breakup
were characterized by means of high-speed photography. The impact pressure was analyzed using
a high resolution piezoelectric senor. The grinding trials showed the impact of the jet characteristics on
the grinding process.
The grinding fluid jet characteristics considered here were divided into three different categories
“coherent jet”, “wave & droplet”, and “atomization”, depending on the grinding fluid nozzle.
The different jet characteristics should already be taken into account when designing a grinding
fluid nozzle. The jet characteristics and breakup have a significant influence on how much of the
workpiece volume can be removed before thermo-mechanical damage occurs. With a higher jet quality,

104
Inventions 2017, 2, 27

namely a breakup that is more “wave & droplet” than “atomization”, a higher specific removed
material volume can be achieved by up to 40%.
Further investigations using high-speed photography are planned to consider the interaction
between the grinding fluid jet and the rotating grinding wheel for different types of jet breakup.
This should deliver a better knowledge of the interaction and the efficient disposal of grinding fluid
on the wheel and within the contact zone leading to a better grinding performance. In addition,
the previous knowledge should be transferred to other grinding processes with higher grinding wheel
circumferential speeds and complex contact conditions.

Acknowledgments: The IGF-research project 18204/N of the FVA—Forschungsvereinigung Antriebstechnik


e.V.—was part of the program Industrielle Gemeinschaftsforschung und—entwicklung (IGF) and was supported
by the German Federal Ministry for Economics and Technology (BMWi) via the Industrial Cooperative Research
Associations (AiF). The authors express their sincere thanks to the collaborating industrial partners for the support
of this research.
Author Contributions: Philip Geilert and André Wagner conceived and designed the experiments; Philip Geilert
performed the experiments and analyzed the data; André Wagner and Carsten Heinzel contributed to the analysis
and the interpretation of results.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Brinksmeier, E.; Heinzel, C.; Meyer, L. Coolant supply conditions and their effect on the workpiece surface
layer in grinding. Prod. Eng. Res. Dev. Ann. WGP 2001, 8, 9–12.
2. Webster, J.A.; Cui, C.; Mindek, R.B., Jr. Grinding fluid application system design. CIRP Ann. 1995, 44, 333–338.
[CrossRef]
3. Rowe, W. 8-Application of Fluids. In Principles of Modern Grinding Technology, 2nd ed.; William Andrew
Publishing: Boston, MA, USA, 2014.
4. Kirsch, B. The impact of contact zone flow rate and bulk cooling on the cooling efficiency in grinding
applying different nozzle designs and grinding wheel textures. CIRP J. Manuf. Sci. Technol. 2017, 18, 179–187.
[CrossRef]
5. Webster, J.; Brinksmeier, E.; Heinzel, C.; Wittmann, M.; Thöns, K. Assessment of grinding fluid effectiveness
in continuous-dress creep feed grinding. CIRP Ann. 2002, 51, 235–240. [CrossRef]
6. Wittmann, M.; Heinzel, C.; Brinksmeier, E. Evaluating the efficiency of coolant supply systems in grinding.
Prod. Eng. Res. Dev. Ann. WGP 2004, 11, 39–42.
7. Meyer, L.; Heinzel, C.; Brinksmeier, E. Analysis and optimization of coolant supply conditions in grinding.
Prod. Eng. Res. Dev. Ann. WGP 2005, 12, 27–30.
8. Cui, C. Experimental Investigation of Thermofluids in the Grinding Zone. Ph.D. Thesis, University of
Connecticut, Fairfield, CT, USA, 1995.
9. Li, D. Encyclopedia of Microfluidics and Nanofluidics; Springer: New York, NY, USA, 2008.
10. Ohnesorge, W.V. Die bildung von tropfen an düsen und die auflösung flüssiger strahlen. J. Appl. Math. Mech.
1936, 16, 355–358. [CrossRef]
11. Oertel, H. Prandtl—Führer Durch die Strömungslehre—Grundlagen und Phänomene, 13th ed.; Springer Vieweg:
Karlsruhe, Germany, 2012.

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

105
inventions
Article
Experimental Analysis for the Use of Sodium
Dodecyl Sulfate as a Soluble Metal Cutting Fluid for
Micromachining with Electroless-Plated Micropencil
Grinding Tools
Peter A. Arrabiyeh 1, *, Martin Bohley 1 , Felix Ströer 2 , Benjamin Kirsch 1 , Jörg Seewig 2
and Jan C. Aurich 1
1 Institute for Manufacturing Technology and Production Systems, University of Kaiserslautern, P.O. Box 3049,
67653 Kaiserslautern, Germany; [email protected] (M.B.); [email protected] (B.K.);
[email protected] (J.C.A.)
2 Institute for Measurement and Sensor-Technology, University of Kaiserslautern, P.O. Box 3049,
67653 Kaiserslautern, Germany; [email protected] (F.S.); [email protected] (J.S.)
* Correspondence: [email protected]; Tel.: +49-631-205-5483

Received: 29 September 2017; Accepted: 8 November 2017; Published: 10 November 2017

Abstract: Microgrinding with micropencil grinding tools (MPGTs) is a flexible and economic process
to machine microstructures in hard and brittle materials. In macrogrinding, cooling and lubrication
are done with metal cutting fluids; their application and influence is well researched. Although it
can be expected that metal cutting fluids also play a decisive role in microgrinding, systematic
investigations can hardly be found. A metal cutting fluid capable of wetting the machining process,
containing quantities as small as 0.02% of the water-soluble fluid sodium dodecyl sulfate was tested in
microgrinding experiments with MPGTs (diameter ~50 μm; abrasive grit size 2–4 μm). The workpiece
material was hardened 16MnCr5.

Keywords: microgrinding; sodium dodecyl sulfate; metal cutting fluid; microstructures; micropencil
grinding tools

1. Introduction
Microcomponents with functional surfaces are becoming an integral part [1] in precision industries
such as biomedicine, aerospace, microelectronics and telecommunications [2]. There is a growing
need for microstructured components, especially ones that manage fluids on the microscopic scale.
Microfluidic components generally use a small sample volume, have a good temperature control and
can downscale analytical equipment for chemical and biomedical analysis [3].
A number of microstructuring processes have been developed to fulfill the market’s need for
microstructured components. Processes like LIGA (lithography, electroplating, and molding), micro
molding and chemical etching techniques are suitable for the mass production of these parts, but lack
the flexibility for small batch production [4]. Micro-end milling [5] and microdrilling are far more
suitable for small batch production lines, but are limited in the hardness of machinable materials [4].
In conventional machining, abrasive processes like grinding use superabrasives made of diamonds
or cBN (cubic boron nitride) grits to machine hard and brittle materials [6]. In the past, miniaturized
versions of these abrasive processes have been developed for microstructuring purposes; among
them microgrinding [1]. Microgrinding has a competitive edge over other microstructuring process,
since it is used as a finishing process that manufactures surfaces with optical quality while minimizing
burr formation [2]. There are two kinds of microgrinding tools in micromachining [4]: thin grinding

Inventions 2017, 2, 29; doi:10.3390/inventions2040029 106 www.mdpi.com/journal/inventions


Inventions 2017, 2, 29

wheels called dicing blades used in the semiconductor industry to cut silicon wafers and produce open
structures and micropencil grinding tools (MPGTs), used for freeform surfaces and microholes [7].
Microgrinding is a process in which the material removal takes place by stochastically distributed
grits [8]. The grits have different protrusions, resulting in material removal at different chip thicknesses.
The value of the chip thickness needs to surpass a minimal size to initiate material removal, otherwise
only elastic and plastic deformation occur during the machining process. Most grits do not reach that
minimal chip thickness, resulting in very high local temperatures due to friction [9]. High temperatures
in turn can cause intense tool wear and a high thermal strain on the workpiece surface that causes
material structure changes and high tensile stresses [9]. A metal cutting fluid (MCF) is needed to
improve the surface quality of machined materials to transport the produced chips away from the
machining zone and to increase tool life [10].
A typical flood supply of MCF is generally not suitable for micromachining, as a high flow
pressure of liquids may influence the tool behavior by applying an additional force [11]. In literature,
new lubrication methods and metal cutting fluids have been developed for micromachining processes.
Brudek et al. compared the roughness values: the arithmetic mean roughness Ra and the mean
roughness depth Rz of micromilled substrates machined with a minimum quantity method to a
process where the workpiece is completely immersed in metal cutting fluid. A variety of commercially
available MCFs as well as a variety of vegetable oils were used. Both lubrication methods showed
similar results with the commercial MCFs having a small edge over the vegetable oils [12]. Nam et al.
used nanofluids in a microdrilling process as an alternative to conventional MCFs. Nanofluids are
composed by a water- or oil-based fluid containing nanoparticles made of materials like graphite,
Al2 O3 , C60 or diamond to increase the thermal conductivity and/or decrease friction in the contact area
via the ball bearing effect [13]. Pham et al. used another alternative MCF by spraying ionic liquids in a
micromilling process while machining aluminum workpieces. Ionic liquids are liquid salts, consisting
of an organic cation and an inorganic anion; they have a low vapor pressure, are non-flammable and
have a high thermal stability. The test series showed that workpieces machined with ionic fluids reach
similar cutting forces and similar roughness values to those machined with commercially available
metal cutting fluids [14]. Overall, the impact of MCF in microgrinding is rarely investigated.
The machine tool used in this paper is located in a clean room and is not capsuled to its
environment; a metal cutting fluid safe for both the user and the machinery is hence needed for
lubrication. Sodium dodecyl sulfate (SDS) is a surfactant and detergent that has lubricating qualities.
It is cheap, causes no health hazards and is even commonly used in the soap and shampoo industry [15].
In this paper, MPGTs with a diameter of ~50 μm and a cBN grit size of 2–4 μm are used to machine
hardened 16MnCr5 steel using distilled water and minimal quantities of the solid surfactant SDS to
lubricate the process. The results are then compared to a dry microgrinding process and one that uses
distilled water only as a coolant. Besides the wear and roughness analysis, the paper offers a tool
characterization pre- and post-machining, a structure characterization as well as a force analysis.

2. Materials and Methods

2.1. Micropencil Grinding Tools


The substrates used for the MPGTs contain a tungsten carbide content of 92%, a cobalt content
of 8% with a grain size of 0.2 μm. The shaft of the substrates has a diameter of 3.175 mm, a bending
strength of 4800 N/mm2 and a Vickers hardness of 1920 ± 50 HV30 (ISO 3878) [16]. Figure 1 depicts
the geometry of a substrate and the two main steps in the manufacturing of MPGTs. A 40◦ cone is
machined on a conventional tool grinding machine onto the substrate to decrease the material removal
in the following precision grinding steps. Using a thin grinding wheel, the cylindrical tip of the
substrate is machined to have a diameter of 44 ± 2 μm at a length of 140 μm. For this paper, a grit
size of 2–4 μm is used. The tool tip diameter must be readjusted when using different grit sizes with
different coating thicknesses to allow the coated tool to reach a diameter of ~50 μm.

107
Inventions 2017, 2, 29

Figure 1. Manufacturing process for micropencil grinding tools (MPGTs): (a) geometry of machined
substrate; (b) microgrinding process for MPGT substrates; (c) electroless plating process for MPGT and
(d) finished MPGT with 2–4 μm grit size.

Following the machining process, the substrate is degreased in an alkaline degreasing solution,
which is then neutralized in a hydrochloric acid solution. Then a thin nickel layer is electroplated
onto the substrate (Figure 2) to provide an active, chemically affine nickel surface. Shrink tubes are
applied to the substrates as a resist to limit the nickel coating to a defined area [17]. Finally, the
electroless-plating process is performed.

Figure 2. Abrasive layer. cBN: cubic boron nitride.

Electroless plating is a process that is based on the principle of ionic reduction. In the solution,
a metal salt, in this case nickel sulfate, provides the solution with Ni2+ free nickel ions. A reducing
agent like sodium hypophosphite can provide the nickel ions with the two missing electrons, slowly
forming a phosphorous nickel coating onto an active surface; the components of the plating solution
are listed in Table 1. The process is suited to manufacture small quantities of MPGTs, with flexible,
custom design choices in regards to its form, diameter, coating thickness, grit size, grit concentration
and grit protrusion [17].
Using the ingredients listed in Table 1, a quantitative energy dispersive X-ray (EDX) analysis
shows that a phosphorous content of 6.01% ± 0.55% can be achieved. A phosphorus content of less than
7% results in a face-centered cubic crystal structure, while an amorphous structure is produced at higher
phosphorus contents [18]. A low phosphorous content generally produces a harder nickel layer [19].

108
Inventions 2017, 2, 29

Table 1. Electroless-plating solution composition [20].

Component Concentration in g/L


Nickel sulfate (NiSO4 ·6H2 O) 30
Sodium hypophosphite (NaH2 PO2 ) 20
Sodium acetate (C2 H3 NaO2 ) 20
Thiourea (CH4 N2 S) 0.0004
Hydrochloric acid (HCl) Adapted to a pH value of 5.2–5.4
cBN grits 4

The abrasive grits are whirled up in the coating solution via a magnetic stirrer. The main
coating time for a monolayered MPGT is 150 s for a grit size of 2–4 μm. After the main coating
time, the magnetic stirrer stops and the grits fall to the bottom of the beaker, allowing to embed the
grits on the MPGT with an additional nickel layer for 90 s (see Figure 2). The final product for a single
layered MPGT can be seen in Figure 1d. Using scanning electron microscopy (SEM) images and an
image processing software, a quantitative analysis was conducted to determine the grit concentration
on the tool. A grit concentration of 35% ± 7% was found.

2.2. Experimental Setup


The results presented in the following chapters were produced on a high precision three-axis
machine tool (Figure 3) mounted on top of a vibration isolated granite plate. The tool spindle is
mounted vertically onto the z-axis on a cross-roller bearing stage. Rotation speeds in the range of
5000–54,000 rpm can be achieved. The X–Y table is guided by air bearings and can move with a
positioning accuracy of <1 μm [5]. A Kistler 3-component dynamometer (9119AA1) dynamometer
for measuring cutting forces is mounted on top of the X–Y table; the workpieces are clamped onto
the dynamometer.

Figure 3. Machine tool for microgrinding and milling. MQL: minimum quantity lubrication.

109
Inventions 2017, 2, 29

A Venturi minimum quantity lubrication (MQL) system is used to spray the machining process
with an air/liquid mixture. By narrowing the cross-section at the nozzle head, a pressure difference
is created through which the liquid is suctioned. The air acts as a transport medium for the liquid.
Flow rates of <100 mL/h are defined as MQL in macro machining. However, considering the small sizes
of tools and structures and the comparably low material removal rates in microgrinding, 100 mL/h
can be defined as flood cooling in microgrinding.
Images of the tools and their respective structures were captured using a scanning electron
microscope (SEM). The structures were analyzed using a confocal microscope (Nanofocus μsurf) with
a 60× magnification lens and a numerical aperture (NA) of 0.9.

2.3. Experimental Procedure


To test the influence of SDS as a metal cutting fluid in the microgrinding process, MPGTs with
diameters ~50 μm were used to machine 500 μm long grooves into hardened 16MnCr5 (SAE5115;
665 HV30 ± 15 HV30 (according to ISO 6507 [21])). The workpiece was face-machined with a larger
pencil grinding tool (diameter = 3.175 mm) to compensate for assembly-related influences and to gain a
flat surface. To test the effect of the soluble lubricant, a small amount of 0.2 g/L was added to a distilled
water medium and was used for the experiments. For comparison, the microgrinding process was
also conducted dry and with distilled water as metal cutting fluid. Both the SDS and distilled water
experiments were conducted with a volume flow rate of 60 ± 10 mL/h and a positive air pressure of
0.65 bar. Both form a fluid film around the tool during the process (Figure 4b); experiments showed
that if the fluid film is interrupted, immediate damage to the abrasive layer occurs.
Based on preliminary studies; a rotation speed of 30,000 rpm (cutting speed of 4.71 m/min)
was applied. Feed rates of 0.05 mm/min and 0.1 mm/min were used at a depth of cut of 5 μm;
Figure 4a shows the parameter combinations studied in this paper with each combination being
repeated three times. The tools were maneuvered to the starting position optically using the camera.
The MPGT is used to scratch the surface of the workpiece to determine the zero position between
tool and workpiece. This results into a positioning accuracy of ±50 μm and hence in an according
deviation of the groove length. The tool rotates in clockwise direction, while the workpiece was given
a feed rate towards the tool.

Figure 4. (a) Microgrinding test series; (b) microgrinding process.

3. Results

3.1. Tool Wear


Figure 5 visualizes the tool wear for four of the parameter combinations (cases) listed in Figure 4a;
the results from parameter combination 2 and 4 were left out of the figure, because they do not differ

110
Inventions 2017, 2, 29

from cases 1 and 3. The tools used in the dry machining process lose their abrasive layer on the face
side of the tool upon entry (Figure 5a). The abrasive layer breaks off the MPGT and rips part of the
layer on the circumference of the tool as well (Figure 5a). It is assumed that high temperatures occur
on the face side of the MPGT due to friction. Tungsten carbide with an 8% cobalt content has a linear
thermal expansion coefficient of 5 × 10−6 –5.2 × 10−6 K−1 which is much smaller than that of nickel
which lies in a range of 12 × 10−6 –13.5 × 10−6 K−1 [22]. Thus, the abrasive layer expands much more
with rising temperature than the substrate, causing the abrasive layer to loosen up until the process
forces eventually result into failure of the abrasive layer.
MPGTs cooled with water showed a slight improvement over the ones used in dry machining,
despite losing the abrasive layer. The tools had a longer tool life than the ones used in dry machining
as can be seen from the abrasive layer breakoff point presented in Figure 5b. The abrasive layer stays
in tact in both cases that use SDS for lubrication (Figure 5c,d); except for one tool used in case six that
broke off 10 μm before finishing its groove. Tool wear is slightly higher for case 6 due to the increase in
feed rate.
An energy-dispersive X-ray analysis (EDX) was performed to determine the iron adhesion on
the face side of the abrasive layer and the nickel adhesion at the bottom of the structure. While the
analysis showed no signs of nickel adhesion in the structure, an iron concentration of 28.65% ± 5.1%
for case 5 and 50.05% ± 0.78% for case 6 was identified. The results are qualitative and require
further investigation; however, a direct relationship between feed rate and material adhesion could be
determined. According to Klocke, adhesions that fill the chip space, increase friction, and therefore the
process temperature and the process forces. Material adhesions increase the wear of the machining
tool by breaking out single grits or even entire grit populations [9].

Figure 5. MPGTs post-microgrinding: (a) case 1; (b) case 3; (c) case 5 and (d) case 6.

111
Inventions 2017, 2, 29

3.2. Structure Analysis


Topographies of the structures machined with all six parameter combinations were measured
using a confocal microscope (Nanofocus μsurf) with a 60× magnification lens and a numerical aperture
of NA = 0.9. Multiple measurements were combined using the stitching algorithms integrated
in the measurement devices software. Missing data points, resulting from the finite numerical
aperture or artefacts on the surfaces, were interpolated using linear interpolation. A first order
plane levelling minimizing the sum of the squared distances was applied to the areal measurement
data. Profiles located in the center of the groove with a length of 400 μm were extracted manually from
the areal data. Calculation of 2D roughness parameters included limitation of the bandwidth using the
Gaussian-filter [23,24] (lc = 80 μm; ls = 0 μm).
To estimate the surface quality, the mean roughness depth Rz and the arithmetical mean roughness
value Ra [25], were calculated according to DIN EN ISO 4288. According to the ISO standard,
five consecutive 80 μm long segments (from a 400 μm long section) were used to calculate the roughness
Rai with i = 1, . . . , 5 of a structure [26], with the first segment starting at the entry point of each groove,
in order to monitor the tool right from the start.
Figure 6a,b display the mean values of Ra and Rz for all six parameter combinations with their
respective standard deviations. A few trends are visible in the diagrams; the first being that grooves
machined with lubrication (Cases 5 and 6), resulted in the smallest roughness values, while grooves
machined dry exhibited overall higher values. The standard deviation of the roughness parameters
is also much smaller for grooves machined with the SDS mixture, which is synonymous for a more
stable process, since less material adhesion on the slot bottom appeared. Another visible trend is that
grooves machined with a feed rate of 0.1 mm/min feature smaller roughness values when compare to
those machined with a 0.05 mm/min feed rate which is a surprising trend considering higher uncut
chip thicknesses at higher feed rates. The lower roughness at higher feeds could be led back to the
higher tool wear and higher material adhesion at larger feed rates.

Figure 6. (a) Roughness values Ra and (b) Roughness values Rz .

The sample structure presented in Figure 7c is one of the structures from case 6. It is chosen to
present some of the more prominent tool-specific and structural characteristics. The first being the

112
Inventions 2017, 2, 29

obvious breakoff at the entry point—this indicates a loss in abrasive grits. Judging from Figure 7a,
many grits at the center of the tool had a rather high protrusion; some of these grits broke off upon
entry due to a lack of grit retention forces at small high grit protrusion and at the same time high uncut
chip thickness, resulting in high loads on those grits. A small 9 μm broad smaller groove in the middle
of the structure (Figure 7c) suggests that the grits at the center of the tool had a higher protrusion.
Figure 7b shows the tool after the machining process and a small circular marking off the pivot of the
tool, which coincidently measures to 9 μm. This pivot is located 2.5 μm off the center of the tool, hence
a reclamping error occurred, producing a step-like structure at the right side of the groove (Figure 7c).
This difference in cutting depths can be led back to a difference in height for parts of the abrasive layer;
this difference in height is marked in Figure 7b but can also be seen in Figure 7a (less grit protrusion).
One final, more common characteristic is the material adhesion to the bottom of the groove, on the up
grinding side.
The manifestation of these characteristics is unique to each individual MPGT. Some of them can
be prevented by sorting out tools that have a higher variation in grit protrusion; however, due to
constant wear, the grits are exposed to during the machining process—it is impossible to prohibit
them completely.
The groove analyzed in in Figure 7c was machined with SDS and a feed rate of 0.1 mm/min.
Different from other structures machined with SDS, the structure has a large material adhesion at the
bottom surface; a characteristic more commonly observed with those machined dry. The structure in
Figure 7c shows almost no signs of burrs and only few, small chippings, while cases 1–4 show much
larger and frequent burr and chipping formations because no material cutting happens after the loss of
an abrasive layer. An example of a dry machined structure is shown in Figure 8.

Figure 7. (a,b) Tool that machined the groove sample, before and post machining and (c) sample of a
structure machined with sodium dodecyl sulfate (SDS) and a feed rate of 0.1 mm/min, measured by
confocal microscopy.

113
Inventions 2017, 2, 29

Figure 8. Sample of a dry machined structure with a feed rate of 0.1 mm/min, measured by confocal
microscopy. A segment is shown with scanning electron microscopy (SEM).

3.3. Force Measurements


The process forces were measured with a Kistler 3-component dynamometer (9119AA1) during
machining. The sampling rate was 10 kHz for all grinding parameters.
After recording, the data was evaluated with National Instruments Software DIAdem. With the
help of a fast Fourier transformation, the actual spindle speed was determined. With this information
a bandpass filter was applied to extract and consider the frequencies in the signal corresponding to
the spindle speed ±20 Hz only. Thus, not the actual process forces were used for this research but the
dynamic process characteristics evaluated and compared. The reason for this configuration is the use
of metal cutting fluid. The impact of the fluid application on the workpiece was higher than the impact
of the cutting process itself.
Figure 9 shows one of the grooves machined for case 4 to demonstrate the correlation between
forces and depth of cut. In Section 1, where the tool is starting to cut the material, the force rates are
rising until the complete tool diameter is in contact. This area is followed by Section 2, where the
cutting conditions seem to be unstable. The cutting forces as well as the feed forces start to rise and
show sharp peaks. At a feed travel of about 150 μm, the forces reach their maximum and some abrasive
grits broke out of the layer or parts of the layer itself broke out, which can be seen in the confocal
image, showing a difference in the resulting geometry of the slot bottom. Until a feed travel of 250 μm
(Section 3), the surface as well as the force amplitudes show an unstable behavior. The depth of cut is
changing in two steps. Until the end of the slot at 500 μm, the force levels as well as the slot bottom
geometry are at a stable cutting regime. At the end of Section 4, the force levels are falling abruptly
after the feed of the machine tool is being stopped.

114
Inventions 2017, 2, 29

Figure 9. Comparison of slot bottom geometry and the equivalent force levels.

4. Conclusions and Outlook


Microgrinding is a process in which a very high amount of rubbing and ploughing occurs.
This can result in high temperatures, high wear of the tools and ultimately a low quality of the
machined structures. This paper presents a systematic investigation on the influence of the application
of metal cutting fluids on tool wear and structure quality.
The application of a new metal cutting fluid (MCF), consisting of small quantities (0.02% in
distilled water) of the solid lubricant sodium dodecyl sulfate (SDS) was tested when microgrinding.
In this test series, microgrinding experiments with monolayered electroless plated micropencil grinding
tools (diameter ~50 μm and grit size 2–4 μm) were conducted to analyze the effect of the new metal
cutting fluid. The results were compared to pure distilled water and dry machining experiments at a
rotational speed of 30,000 rpm and the two-feed rate variations of 0.05 mm/min and 0.1 mm/min.
The results showed that the micropencil grinding tools (MPGTs) used in dry machining and with
distilled water were unable to complete a 500-μm long groove without losing part of their abrasive
layer. In contrast, the grooves machined with SDS were completed with the tool intact.
The structures were analyzed using a confocal microscope. Both the roughness values and the
structure characteristics were examined. Highest roughness values were measured for the dry grinding
cases and lowest roughness values for cases machined with SDS. Besides the lower roughness, a more
stable process was achieved with the application of SDS. This was manifested by small standard

115
Inventions 2017, 2, 29

deviations of the roughness values. The application of SDS also resulted into less burr formation and
chipping, as well as less adhesions on the bottom surface.
In conclusion, the results revealed that the process stability and the quality of the machined
structures is highly influenced by the application of metal cutting fluids. Future work will hence deal
with the investigation of the flow rate and the SDS concentration. In addition, tool and machining
parameter case studies are required to explore the possibilities and limitations of increasing both the
tool life and the productivity of the process. A further analysis of structure and tool characteristics
will be conducted to get a better understanding of the tool wear characteristics and their influence on
the structures.

Acknowledgments: This research was funded by the German Research Foundation (DFG) within the Collaborative
Research Center 926 “Microscale Morphology of Component Surfaces”—through the subproject B09 “Geometrical
Structuring of Component Surfaces by Microgrinding”.
Author Contributions: Peter A. Arrabiyeh has conceived, designed, performed and documented the microgrinding
and electroless plating experiments. Peter A. Arrabiyeh, Martin Bohley and Felix Ströer conducted the structural
analysis of machined microstructures. Martin Bohley analyzed the data collected from the dynamometer.
Peter A. Arrabiyeh conducted SEM and EDX analysis. Benjamin Kirsch supervised the present study and
helped to discuss and analyze the results. Jan C. Aurich initiated the study. All authors were highly involved in
writing the paper.
Conflicts of Interest: The authors declare no conflicts of interest.

References
1. Feng, J. Microgrinding of Ceramic Materials. Ph.D. Thesis, Michigan Technological University,
Houghton, MI, USA, 2010.
2. Pratap, A.; Patra, K.; Dyakonov, A.A. Manufacturing miniature products by micro grinding: A review.
Procedia Eng. 2016, 150, 969–974. [CrossRef]
3. Wensink, H. Fabrication of Microstructures by Powder Blasting. Ph.D. Thesis, University of Twente,
Enschede, The Netherlands, 2002.
4. Engmann, J. Galvanisch Gebundene Mikroschleifstifte. Entwicklung, Herstellung und Einsatz, Als Ms. Gedr;
Produktionstechnische Berichte aus dem FBK 01/2011; Technischen Universität Kaiserslautern:
Kaiserslautern, Germany, 2011.
5. Reichenbach, I.G.; Aurich, J.C. Untersuchung der oberflächengüte beim mikrofräsen—einfluss
von prozessparametern und einstellwinkel der nebenschneide bei 48 μm mikroschaftwerkzeugen.
Wt Werkstattstechnik Online 2013, 103/11–12, 847–852.
6. Brinksmeier, E.; Mutlugünes, Y.; Klocke, F.; Aurich, J.C.; Shore, P.; Ohmori, H. Ultra precision grinding.
CIRP Ann. Manuf. Technol. 2010, 59, 652–671. [CrossRef]
7. Hoffmeister, H.-W.; Hlavac, M. Schleifen von Mikrostrukturen. In Tagungsband des 10. Feinbearbeitungskolloqiums
in Braunschweig; Vulkan-Verlag: Essen, Germany, 2002; pp. 7.1–7.24.
8. Setti, D.; Sinha, M.K.; Ghosh, S.; Rao, P.V. Performance evaluation of Ti–6Al–4V grinding using chip formation
and coefficient of friction under the influence of nanofluids. Int. J. Mach. Tools Manuf. 2015, 88, 237–248.
[CrossRef]
9. Klocke, F. Manufacturing Processes 2: Grinding, Honing, Lapping; Springer: Berlin, Germany, 2009.
10. Brinksmeier, E.; Heinzel, C.; Wittmann, M. Friction, cooling and lubrication in grinding. CIRP Ann.
Manuf. Technol. 1999, 48, 581–598. [CrossRef]
11. Dornfeld, D.; Min, S.; Takeuchi, Y. Recent advances in mechanical micro-machinability of copper 101 using
tungsten carbide micro-end mills. Int. J. Mach. Tools Manuf. 2006, 55, 745–768.
12. Brudek, G. Beiträge zur Prozessanalyse in der Mikrozerspanung. Insbesondere für das Mikrofräsen; Berichte aus
dem Institut für Konstruktions- und Fertigungstechnik Bd. 8; Shaker: Aachen, Germany, 2007.
13. Nam, J.S.; Kim, D.H.; Chung, H.; Lee, S.W. Optimization of environmentally benign micro-drilling process
with nanofluid minimum quantity lubrication using response surface methodology and genetic algorithm.
J. Clean. Prod. 2015, 102, 428–436. [CrossRef]
14. Pham, M.Q.; Yoon, H.S.; Khare, V.; Ahn, S.H. Evaluation of ionic liquids as lubricants in micro milling-process
capability and sustainability. J. Clean. Prod. 2014, 76, 167–173. [CrossRef]

116
Inventions 2017, 2, 29

15. Smulders, E.; Rybinski, W.-V.; Sung, E.; Rähse, W.; Steber, J.; Wiebel, F. Nordskog: Ullmann’s Encyclopedia of
Industrial Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2000.
16. DIN EN ISO 3878, Hardmetals, Vickers Hardness Test, Identical with ISO 3878:1983; International Organization
for Standardization: Geneva, Switzerland, 1991.
17. Arrabiyeh, P.A.; Kirsch, B.; Aurich, J.C. Development of micro pencil grinding tools via an electroless plating
process. J. Micro Nano Manuf. 2017, 5, 011002. [CrossRef]
18. Lin, K.L.; Hwang, J.W. Effect of thiourea and lead acetate on the deposition of electroless nickel.
Mater. Chem. Phys. 2002, 76, 204–211. [CrossRef]
19. Mallory, G.O.; Hajdu, J.B. Electroless Plating: Fundamentals and Applications; American Electroplaters & Surface
Finishers Society: Orlando, FL, USA, 1990.
20. Kirsch, B.; Bohley, M.; Arrabiyeh, P.A.; Aurich, J.C. Application of ultra-small micro grinding and micro
milling tools: Possibilities and limitations. Micromachines 2017, 8, 261. [CrossRef]
21. DIN EN ISO 6507-1:2006-03 Metallic Materials—Vickers Hardness Test—Part 1: Test Method; International
Organization for Standardization: Geneva, Switzerland, 2006.
22. Granta Design Ltd. CES EduPack Software; Granta Design Ltd.: Cambridge, UK, 2012.
23. DIN EN ISO 16610-21:2013-06 Geometrical Product Specifications (GPS)—Filtration—Part 21: Linear Profile
Filters: Gaussian Filters; International Organization for Standardization: Geneva, Switzerland, 2012.
24. DIN EN ISO 11562:1998-09 Geometrical Product Specifications (GPS)—Surface Texture: Profile Method—Metrological
Characteristics of Phase Correct Filters; International Organization for Standardization: Geneva, Switzerland, 1997.
25. DIN EN ISO 4287:2010-07 Geometrical Product Specifications (GPS)—Surface Texture: Profile
Method—Terms, Definitions and Surface Texture Parameters; International Organization for Standardization:
Geneva, Switzerland, 2009.
26. DIN EN ISO 4288:1998-08 Geometrical Product Specifications (GPS)—Surface Texture: Profile Method—Rules
and Procedures for the Assessment of Surface Texture; International Organization for Standardization:
Geneva, Switzerland, 1997.

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

117
inventions
Article
Model Development for Optimum Setup Conditions
that Satisfy Three Stability Criteria of Centerless
Grinding Systems
Fukuo Hashimoto
Advanced Finishing Technology Ltd., Akron, OH 44319, USA; [email protected]

Received: 5 September 2017; Accepted: 16 September 2017; Published: 21 September 2017

Abstract: The centerless grinding process demonstrates superior grinding accuracy with extremely
high productivity, but only if the setup conditions are properly set up. Otherwise, various unfavorable
phenomena manifest during the grinding processes and become serious obstacles to achieving that
high quality and productivity. These phenomena are associated with the fundamental stabilities of
the centerless grinding system, so it is essential to keep the system stable by setting up the appropriate
grinding conditions. This paper describes the development of a model for finding the setup conditions
that simultaneously satisfy the three stability criteria of centerless grinding systems: (1) work rotation
stability for safe operations; (2) geometrical rounding stability for better roundness; and (3) dynamic
system stability for chatter-free grinding. The objective of the model development is to produce
combinations of optimal setup conditions as the outputs of the model, and to rank the priority
of the outputs using PI (performance index) functions based on the process aims (productivity or
accuracy). The paper demonstrates that the developed model, named Opt-Setup Master, can generate
the optimum setup conditions to ensure safe operations, better roundness and chatter-free grinding.
It provides practical setup conditions as well as scientific parameters and fundamental grinding
parameters. Finally, the paper verifies that the Opt-Setup Master provides the setup conditions that
simultaneously satisfy all three stability criteria of the centerless grinding system.

Keywords: grinding; centerless grinding; process optimization; safe operation; quality; productivity

1. Introduction
The centerless grinding method has been extensively applied for the production of cylindrical
components such as rings, rollers, and pins. It is estimated that a single car has more than 2000 parts
finished by centerless grinding processes. The centerless grinding process demonstrates extremely high
productivity with very high grinding accuracy in OD size, roundness and surface integrity. However,
its superior performance compared to other grinding methods can be achieved only if the grinding
conditions are properly set up; otherwise, various unfavorable phenomena, such as slippages in work
rotation, deformed roundness and chatter vibrations, appear during the grinding process and lead to
deterioration in grinding performance [1].
These huge advantages and disadvantages come from the unique work-holding features of the
centerless grinding system: (1) a loose hold on the workpiece without any mechanical constraints;
(2) the work friction brake/drive mechanism of the work rotation; and (3) a self-centering mechanism
called “regenerative centering”. The grinding process is very sensitive to these unique centerless setup
conditions, so it is essential to secure the grinding system’s stability by setting it up appropriately.
This requires controlling the three fundamental stability issues caused by the work-holding features of
centerless grinding. These are: (1) work rotation stability; (2) geometrical rounding stability; and (3)
dynamic system stability.

Inventions 2017, 2, 26; doi:10.3390/inventions2040026 118 www.mdpi.com/journal/inventions


Inventions 2017, 2, 26

Work rotation stability is related to the work friction brake/drive mechanism of the centerless
grinding system. The regulating wheel is in rolling-sliding contact with the workpiece, and provides
the friction force to the workpiece that drives or brakes the work rotation. In this unique mechanism,
the work rotates with almost the same peripheral velocity of the regulating wheel during the stable
grinding process, in which the torque created by the grinding force balances with the torque from the
friction forces acting on the regulating wheel and the blade top surface. However, under heavy grinding
with excessive grinding force, control over the work rotation speed is lost due to the broken torque
equilibrium, and it increases toward the grinding wheel speed. This phenomenon, called “spinners”,
can cause dangerous accidents and should be avoided in order to maintain safe operations. The author
is a pioneer of the study of work rotation stability and has shown that there exists an absolute safe zone
where spinners do not develop [2]. The setup guidelines for safe operations are well established in the
literature, and the means of satisfying the work rotation stability criterion have been demonstrated [3].
Geometrical rounding stability is related to the work-holding conditions and regenerative
centering effects. Although centerless grinding technology has been around for 100 years since
the method was patented by L.R. Heim in 1917 [1], a great deal of effort was exerted by early research
pioneers to understand its rounding mechanism, and significant papers have been published [4–9].
The theory of the rounding mechanism has been well established, the setup guidelines for achieving
better roundness have been described, and the means of satisfying the stability criteria have been
clarified. The stability criteria assume that the grinding system consists of solid bodies and is
dynamically stable. Under certain work-holding conditions, a specific number of lobes on the
roundness of the workpiece appear or cannot be removed. It is crucial to minimize roundness
errors by selecting the proper setup conditions.
Dynamic system stability is related to the work-regenerative chatter vibration caused by the
instability of the centerless grinding system, including the machine dynamics. The chatter vibration in
centerless grinding is very severe and builds up very fast. In general, the amplitude growth rate is 10 to
100 times greater than that of center-type grinding processes, and is caused by the wheel-regenerative
chatter vibration. Significant investigations have been carried out by many researchers [10–13] to
understand dynamic system stability and suppress the chatter vibration. The system stability criterion
has been well established, and the setup condition guidelines for chatter-free grinding are available in
the literature.
As mentioned above, the setup guidelines for satisfying each stable criterion have been established.
However, the setup operations of centerless grinding still rely on experimental skill and the
trial-and-error method. Even though each stable criterion can be satisfied individually by carefully
choosing the setup conditions, it is almost impossible to simultaneously satisfy all three of the centerless
grinding system’s stability criteria. Therefore, a special analytical tool for finding the optimum
combination of setup conditions is greatly needed [14].
The objective of this paper is to describe the development of an analytical model capable of finding
the optimum combination of setup conditions that satisfies all three stability criteria at the same time.
This paper describes the structure of the developed model, which consists of the input-information
session, the data bank that stores all the parameters required for the model calculations, the PI
(performance index) functions for assessing the setup conditions based on the process aim (productivity
or accuracy), and the output-information session.
Further, this paper explains the algorithm of the model and shows how to find the setup conditions
that meet the three stability criteria simultaneously. The developed model, named Opt-Setup Master,
is verified through case studies in which workpieces with various sizes are ground with different
grinding machines. Finally, the Opt-Setup Master demonstrates its capability to generate the optimum
setup conditions that satisfy all three stability criteria, and to provide the grinding conditions that will
provide safe grinding operations and chatter-free grinding with improved grinding accuracy.

119
Inventions 2017, 2, 26

2. Basic Setup Conditions in Centerless Grinding


The basic setup parameters in centerless grinding are the blade angle θ, the center height angle γ
and the work rotational speed nw , as shown in Figure 1.

Figure 1. Setup conditions in centerless grinding.

The set (θ, γ, nw ) of these parameters is called the “setup condition” in this paper, and it
significantly affects centerless grinding performance. In practice, the work center height CH (instead
of γ) and the RW (regulating wheel) rotation speed Nr (rpm) are used because these parameters can be
directly set up on the machine. The center height CH (mm) has the following relationship with the
center height angle γ (◦ ) when angles α and β are small.

γ = α+β (1)

2CH ∼ 2CH
α = sin−1 = (2)
( Dr + Dw ) ( Dr + Dw )
2CH 2CH
β = sin−1   ∼
=   (3)
D g + Dw D g + Dw
 
3.14 Dg + Dw ( Dr + Dw ) ◦
CH (mm) =   γ( ) (4)
360 Dg + Dr + 2Dw
The work rotation speed nw is controlled by the RW friction drive/brake mechanism. Figure 2
shows test results of normal grinding force Fn, the friction coefficient μr and the rolling-sliding velocity
between RW and the workpiece during an infeed centerless grinding process [15]. In steady state
grinding, the sliding velocity ΔV, defined as (Vw − Vr ), is about +0.008 m/s, and the slippage ratio
ΔV/Vr is about 2%, where Vw and Vr are the work and RW peripheral velocities, respectively. Since the
sliding velocity is very small, the work rotation speed nw can be represented by:

Dr Nr (rpm)
nw (rps) ∼
= (5)
60Dw

In the model development, the scientific parameters (θ, γ, nw ) are used for the analysis of the
optimum setup condition, and the practical parameters (θ, CH, Nr ) are the outputs of the model.

120
Inventions 2017, 2, 26

Figure 2. Infeed centerless grinding process. Dr = 255 mm, Dw = 30 mm, Nr = 30 rpm, Vw = 0.4 m/s,
Sliding velocity: 0.008 m/s, Slippage ratio: 2%.

3. Centerless Grinding Systems and the Characteristic Equation


Since the three stability criteria influence each other and are significantly affected by the setup
conditions, it is necessary to assess these stabilities as a total system—including the machine dynamic
characteristics, the centerless grinding mechanism and the grinding processes. Figure 3 shows
a block diagram of the centerless grinding system. The system consists of the regenerative centering
mechanism [16], the regenerative function [17], the relationship between depth-of-cut and the normal
grinding force, the contact stiffness of the wheels, the wheel filtering functions, and the machine
dynamics [12]. The dynamic behavior of the rounding mechanism can be investigated based on the
characteristic equation of the closed loop centerless grinding system in Figure 3.

Regenerative Regenerative Grinding where,


Slide center-function function stiffness Normal ɸʚ= sin(ઞ)/cos(ɽ-ɴ)
feed + 1 force (1-ɸ) = cos(ɽ-ɴ)/cos(ɽ-ɲ)
f z cs 1 − e −2πS b ⋅ kw′ Fn s = ʍ + jn (s: Laplace operator)
1 − ε ′e −ϕ1S + zcr (1 − ε )e −ϕ 2 S
- ੮1: Blade phase angle
Contact Compliance of regulating wheel ੮2: Regulating wheel phase angle
+ (1 − ε ) b: Grinding width
zcr b ⋅ kcr′ kʚw: Specific grinding stiffness
+
Contact Compliance of grinding wheel kʚcr: Contact stiffness of regulating wheel
+ 1 kʚcs : Contact stiffness of grinding wheel
b ⋅ kcs′ ʆ: Mode number
+ Uʆ: ʆ-th. mode orientation factor
Compliance of grinding machine
uν kmʆ: ʆ-th. mode machine static stiffness
¦
ν k mν
Gν (s) G(s): Machine dynamic transfer function
Zcs: Grinding wheel filter function
Zcr: Regulating wheel filter function

Figure 3. Block diagram of centerless grinding system.

The characteristic function is represented by:

1
− = g(s) (6)
f (s)

where  
1 Zcs 1 − e−2πs
− =− (7)
f (s) 1 − ε e− ϕ1 s + Zcr (1 − ε)e− ϕ2 s

1 (1 − ε ) 1
g(s) = bk w  + + Gm (s) (8)
bkcs bkcr km

121
Inventions 2017, 2, 26

By solving the characteristic roots of Equation (6), the dynamic rounding stability can be evaluated
and the transient behavior of the waviness amplitude in work roundness can be calculated during the
grinding process. The characteristic root can be represented by:

s = σ + jn (9)

where s is the Laplace operator, σ is the amplitude growth rate per unit radian, and n is the number
of lobes in the work roundness. The transient of the amplitude change A(t) on roundness waviness
during the grinding process can be expressed by:

A(t) = A0 exp(2πnw σt) (10)

where A0 is the initial amplitude of the waviness, nw is the work rotation speed in rps and t is the
grinding time. When σ is positive, the amplitude of n lobes grows with grinding time t and the
grinding process can be identified as the chatter vibration. In case of σ < 0, the amplitude of n lobes is
decreased with grinding time t, and the grinding process becomes stable with improved roundness.
When the effect of machine vibration is negligible, the response of the transfer function Gm(s) is
degenerated to a constant and the resulting system is of a kinematic nature, referred to as “geometric
rounding stability” [12]. Then, the characteristic equation is simplified as:
  
1 − e−2πs 1 (1 − ε ) b
−  − −
= kw   + + (11)
1−ε e ϕ 1 s + (1 − ε ) e ϕ 2 s k cs kcr km

4. Three Stability Criteria in Centerless Grinding

4.1. Work Rotation Stability Criterion


Figure 4 shows the torques acing on the workpiece during the centerless grinding process. Tg is
the grinding torque given by the tangential grinding force Ft. Tb and Tr are the friction torques acting
on the blade and the regulating wheel, respectively. Under the stable grinding process, the following
torque-quilibrium relationship is maintained. The work peripheral velocity Vw is controlled by the
friction drive/brake mechanism of RW, and becomes almost the same as the RW peripheral velocity Vr.

Stable grinding : Tg = Tb + Tr, Vw ∼


= Vr (12)

Vg Vw Vr
Regulating wheel
Grinding wheel

Workpiece

Tb

Tg
Tr
Blade
Figure 4. Torques acting on workpiece during centerless grinding.

However, once this quilibrium condition is broken by the excessive grinding torque Tg overcoming
the friction torques (Tb + Tr) during grinding, the work velocity Vw suddenly increases toward the
grinding wheel speed Vg.

Unstable grinding : Tg > Tb + Tr, Vw ∼


= Vg (13)

122
Inventions 2017, 2, 26

This phenomenon, called “spinners”, can create a potentially very dangerous situation and should
be avoided for safe operations.
Figure 5a,b show the geometrical arrangement of the centerless grinding process and the
forces acting on the workpiece at any cut section perpendicular to the work axis l during grinding.
The variables fT and fN represent the tangential and normal grinding forces per unit width at the cut
section. Rb and Rr are the resultant forces, while μb and μr are the friction coefficients at the contact
points with the blade and the RW, respectively. w(l) is the work weight per unit width at the cut
Section 1. The torque equilibrium equation can be written by:

L
dω ( B1 μr (l ) + B2 ) f T (l ) − (C1 μr (l ) + C2 )w(l )
I = r (l ) dl (14)
dt ( A1 μr ( l ) + A2 )
0

where
A1 = μb cos(θ − α) − sin(θ − α) (15)

A2 = μb sin(θ − α) + cos(θ − α) (16)

B1 = A1 − μb (sin γ + k cos γ) − [(1 + kμb ) sin(θ + β) + (k − μb ) cos(θ + β)] (17)

B2 = A2 − μb (cos γ − k sin γ) (18)

C1 = sin θ − μb (cos θ − sin α) (19)

C2 = μb cos α (20)

I and ω are the mass moment of inertia and the angular velocity of the workpiece. k is the force
ratio (fN /fT ). For convenience, the plus sign of μr is assigned to the downward friction force and the
minus sign is assigned to the upward one.

(a) (b)

Figure 5. Arrangement of centerless grinding and forces acting on workpiece. (a) Forces acting on
workpiece during centerless grinding; (b) Cylindrical workpiece.

The generalized motion Equation (14) is applicable to any cylindrical-shaped workpiece; for
example, simple cylindrical, tapered, and multiple stepped diameter workpieces. Equation (14)
indicates that, in addition to being affected by the primary setup conditions (θ, γ), the rotational
motion of the workpiece is affected by the grinding forces and the friction force on RW.
The upper-limit tangential grinding force fU under the stable grinding condition is derived from
Equation (14).
(C1 μr0 + C2 )
fU = w (21)
( B1 μr0 + B2 )

123
Inventions 2017, 2, 26

where μr0 is the maximum static friction coefficient of RW. When the tangential grinding force fT is
smaller than fU , the work rotation speed Vw can be controlled with the RW speed Vr.
Figure 6 shows the results of the calculation of fU with respect to the blade angle θ with various
friction coefficients μr0 . The grinding force fU is normalized with the diameter d of a simple cylindrical
workpiece made of steel. fU increases with increased θ. When θ is greater than a certain angle with
μr0 , the fU value becomes infinite. Under this condition, there is no risk of the spinners phenomenon
occurring. The zone with the infinite fU value is called the “safe operation zone.” For instance, there is
no limit on fU when a blade of θ > 42◦ is used with an RW of μr0 = 0.25.

Figure 6. Upper limit tangential grinding force.

Figure 7 shows the safe operation zones under various μr0 values on the (θ–γ) chart and provides
guidelines for satisfying the work rotation stability criterion (WRSC). Stable grinding without any risk
of spinners can be obtained by selecting the set of (θ, γ) from the safe operation zone, and the WRSC is
satisfied with the setup conditions (θ, γ).

Figure 7. Safe operation chart.

4.2. Geometrical Rounding Stability Criterion


When the grinding system is stable and the influence of the machine dynamics on the rounding
mechanism is negligible, the stability of the rounding mechanism is predominantly affected by the

124
Inventions 2017, 2, 26

geometrical arrangement of the centerless grinding system under the solid-body machine structure.
By analyzing the characteristic roots of Equation (11), the effect of the center height angle γ on
geometrical rounding stability can be assessed. Figure 8a shows the characteristic root distributions
for the odd lobes and the even lobes [14]. When a lower center height angle such as γ < 3◦ is set up,
the amplitude growth rates of the 3, 5, and 7 lobes become close to zero, and the roundness error due
to these odd lobes cannot be improved during the grinding process. On the other hand, when a higher
center height angle such as γ > 9◦ is set up, the amplitude growth rates of even lobes like 18, 20, and 22
become close to zero and the roundness error cannot be improved due to the even-lobe waviness.

10

Conditions
Amplitude growth rate ʍ

8
Blade angle ɽ=30°

Roundness ʅm
Stock removed in dia. =35-40ʅm
ʅ̉
4 6
Work size (dia. x width)
X: 9 x 30 mm
Roundness
4
ȴ: 14 x 30 mm
2
o: 20 x 30 mm
2

ઞopt ˕ 0
ˏ 0
Center-height angle ߛι 5 10 0 2 4 6 8 10
Center-height angle ߛι Center-height angle ߛι
(a) (b) (c)

Figure 8. Geometrical rounding stability. (a) Center-height angle vs. amplitude growth rates [14];
(b) Effect of γ on roundness [6]; (c) Effect of γ on roundness [16].

These results suggest the existence of an optimum center height angle γopt that yields a minimum
roundness error, and γopt = 6.7◦ has been proposed as that optimum angle [12]. Miyashita et al.
reported the experimental results shown in Figure 8b, and indicated that the optimum center height
angle exists around 7◦ [6]. Rowe et al. reported on theoretical and experimental analysis of the
rounding mechanism of a workpiece with a flat face. Figure 8c shows the effect of the center height
angle on the roundness error. The grinding test results verified that the optimum center height angle
γopt exists around 6◦ –8◦ [8].
It is well known that an odd number of lobes appears under a lower center height condition such
as γ < 3◦ , as shown in Figure 9a. To minimize the roundness error with odd numbers of lobes, it is
recommended that lower center height angles be avoided as much as possible. Under a relatively higher
center height condition, a specific even number of lobes appears, as shown in Figure 9b. Where the center
height angle is known, the even number of lobes ne that will appear can be found by 180/γ [6].

⹄ࡺ⹕⸣ 䃯ᮤ䓺
RW 䃯ᮤ䓺
GW ⹄ࡺ⹕⸣
GW RW
Ȗ
ਇᶯ
ਇᶯ

Blade Blade

7 lobes Ȗ˙5º Ȗ˙7.5º Ȗ˙9º


3 lobes 5 lobes 36 lobes 24 lobes 20 lobes

ne=180/5=36 ne=180/7.5=24 ne=180/9=20

(a) (b)

Figure 9. Effect of center-height angle on rounding mechanism. (a) Odd lobe appearance at lower γ;
(b) Even lobe appearance at a specific γ.

125
Inventions 2017, 2, 26

These even numbers of lobes are the result of the geometrical rounding instability caused by the
geometrical arrangement of the regulating wheel or the blade. Figures 10 and 11 show the geometrical
rounding stability criteria of the regulating wheel and the blade, respectively.
Figure 10a is an example of regulating wheel geometrical rounding instability. When a peak
of waviness on the work roundness contacts with the regulating wheel and the waviness becomes
a valley at the grinding point (and vice versa), the waviness error cannot be removed during the
grinding process. Conversely, Figure 10b is an example of regulating wheel geometrical stability.
When a peak contacts with the regulating wheel and the waviness becomes a peak at the grinding
point (and vice versa), the waviness error can be removed during the process. Therefore, the regulating
wheel geometrical rounding stability criterion (RW − GRSC) can be summarized as follows:

180
( RW − GRSC ) Unstable : = Even integer (22)
γ

180
( RW − GRSC ) Stable : = Odd integer (23)
γ

GW RW GW RW
n=16 n=16
੘ ੘

[180/੘]=16: Even
UnstableEven [180°/੘°]=Even integer
[180/੘]=16:

(a)
GW RW
GW RW
n=15 n=15
੘ ੘

Stable Odd [180°/੘°]=Odd[180/੘]=15:


[180/੘]=15: integer Odd

(b)

Figure 10. Regulating wheel (RW) geometrical rounding stability criteria. (a) RW geometrical rounding
unstable conditions; (b) RW geometrical rounding stable conditions.
Likewise, Figure 11a is an example of blade geometrical rounding instability. When a peak of
waviness contacts with the top surface of the blade and the waviness becomes a peak at the grinding
point (and vice versa), the waviness error cannot be removed. In addition, again, Figure 11b is
an example of blade geometrical rounding stability. When a valley contacts with the top surface
of the blade and the waviness becomes a peak at the grinding point (and vice versa), the waviness
error can be removed. Similarly, the blade geometrical rounding stability criterion (B − GRSC) can be
summarized as follows:
90 − θ − β
( B − GRSC ) Unstable : = Even integer (24)
γ

90 − θ − β
( B − GRSC ) Stable : = Odd integer (25)
γ

126
Inventions 2017, 2, 26

GW n=16 GW n=16
ɴ ɴ

ɽ ɽ

Unstable [(90°-ɽ°-ɴ°)/੘°]=Even integer

(a)
GW GW
n=15 n=15
ɴ ɴ

ɽ ɽ

Stable [(90°-ɽ°-ɴ°)/੘°]=Odd integer

(b)

Figure 11. Blade geometrical rounding stability criteria. (a) Blade geometrical rounding unstable
conditions; (b) Blade geometrical rounding stable conditions.

To satisfy the RW geometrical rounding stability criterion (RW − GRSC), the center height angle
γ is determined by:
180
RW − GRSC Stable : γ = (26)
Iodd
where Iodd is an odd integer. Also, the blade geometrical stability criterion (B − GRSC) can be satisfied
by setting up the center height angle γ that can be calculated by the following equation.

(90 − θ )
B − GRSC Stable : γ =   (27)
( Dr + Dw)
Iodd + ( Dg+ Dr +2Dw)

4.3. Dynamic System Stability Criterion


As mentioned, the chatter vibration in the centerless grinding system is very severe and builds
up fast because the work-regenerative self-excited vibration has such a high amplitude growth rate,
as shown in Figure 12a. This phenomenon not only deteriorates grinding accuracy and productivity,
but also threatens safe operations. To achieve a stable grinding process with high grinding accuracy as
shown in Figure 12b, satisfying the dynamic system stability criterion is imperative.
Study of the characteristic root distributions of Equation (6) clarifies the generation mechanism of
the chatter vibration [12], and the chatter generation zones are revealed on the (n·γ–n·nw ) diagram for
the dynamic system stability criterion (DSSC). Figure 13 plots the 3D positive growth rates σ of the
characteristic roots on the (n·γ–n·nw ) diagram. The chatter zones are shown as “mountains” located
near the natural frequencies in the (n·nw ) axis. The higher the height of the mountain, the more severe
the chatter vibration. Since the chatter mountains are in 0 < (n·γ) < 180◦ , they generate even numbers
of lobes during chatter vibration.
The chatter zones located in 180◦ < (n·γ) < 360◦ generate odd numbers of lobes and generate
chatter vibration only under higher γ values. A straight line (nw /γ) through the origin is determined
when the center height angle γ and the work speed nw are given. Where the straight line hits
a mountain, chatter vibration occurs at that frequency. In other words, where the straight line given
by the ratio (nw /γ) does not hit any mountains, the DSSC is satisfied and chatter-free grinding is
achieved. Figure 13 shows distinct areas that satisfy the DSSC and gives the ranges of (nw /γ) that
provide chatter-free grinding.

127
Inventions 2017, 2, 26

Out-of-roundness of
ground workpiece
- Mag. x 5,000
- Filter 1-50

0.2 ʅm
n=12 lobes Dynamic component of grinding force

nw=8.9 rps, ઞ=9°, nw=2.1 rps, ઞ=9°,


nw/ઞ=0.99 nw/ઞ=0.23
(a) (b)

Figure 12. Dynamic stability of centerless grinding system (experiment). (a) Unstable grinding process
(chatter); (b) Stable grinding process.

Figure 13. Dynamic system stability diagram for Machine A. Conditions: b = 70 mm, k’w = 2 kN/mm·mm,
k’cr = 0.3 kN/mm·mm, k’cs = 1 kN/mm·mm, kmr = 0.1 kN/μm, kms = 0.15 kN/μm, km1 = 0.3 kN/μm,
fnr = 100 Hz, fns = 200 Hz, fn1 = 150 Hz, ζr = 0.05, ζs = 0.05, ζ1 = 0.05.

Each grinding machine has its own dynamic characteristics with different natural frequencies,
and a (n·γ–n·nw ) diagram can be plotted that shows its unique chatter zones. The chatter zones are
identified by conducting systematic grinding tests. Figure 14a shows the chatter zones of grinding
machine A. The chatter zones located near the natural frequencies of 100 Hz, 150 Hz and 200 Hz are
shown on the vertical axis (n·nw ) Hz. These chatter zones are located in 0 < (n·γ) < 180◦ and generate
even numbers of lobes, while the zones located in 180◦ < (n·γ) < 360◦ generate odd numbers of lobes
where γ > 6.7◦ . Figure 14b shows the chatter zones for grinding machine B. Machine B is designed with
high, rigid structures and possesses spindles with very high stiffness. The first chatter zone appears at
the natural frequency of 430 Hz, which is very high, and machine B creates wider chatter-free regions
than conventional machine A.
Figure 15a shows the chatter zones of grinding machine A plotted on a (γ–nw ) chart. The practical
(γ–nw ) chart can describe chatter zones, but cannot describe the chatter generation mechanism,
the chatter zones’ various vibration modes, or the areas where stable grinding can occur. However,

128
Inventions 2017, 2, 26

the (γ–nw ) chart is very effective in setup operations when used along with the analytical (n·γ–n·nw )
diagram. The ranges of chatter zones in (nw /γ) are explicitly given, as shown in Figure 15b.

250 600 Chatter


(lobes)x(work speed) n·nw Hz

(lobes)x(work speed) n·nw Hz


zone
200
Chatter
zone 400
150

100
200
Work size
50 O dia. 10 x L 70 mm work
Worksize
size
•dia. 300 x L 70 mm •dia. 30 x LL70
Dia.30 70 mm
mm

0 90 180 270 360 0 90 180 270 360


(lobes)x(center-height angle) n·ઞ° (lobes)x(center-height angle) n·ઞ°
(a) (b)

Figure 14. Chatter zones on (n·γ–n·nw ) diagram. (a) Grinding machine A; (b) Grinding machine B.

No chatter 100Hz chatter 150Hz chatter 200Hz chatter


[nw γ ]H
20
15
Work speed nw rps.

Work speed nw rps.

15
Chatter region
10
10
ª nw º
«γ »
¬ ¼ L1 5
5 ª nw º
«γ »
¬ ¼ L2

0 3 6 9 0 3 6 9 12
Center-height angle ઞ ° Center-height angle ઞ °
(a) (b)

Figure 15. Dynamic system stability criterion for machine A. (a) Chatter zones confirmed by grinding
tests; (b) Dynamic system stability criterion.

Three types of stable, no-chatter zones exist for high work speed and low work speed regions.
The chatter-free conditions for machine A are:
1. (nw /γ) H > 3.0 (high-speed chatter-free zone; KH)
2. (nw /γ) L1 < 0.6 when γ is lower (low-speed chatter-free zone 1; KL1)
3. (nw /γ) L2 < 0.28 when γ is higher (low-speed chatter-free zone 2; KL2)
Figure 16a,b redraw Figure 14b to include chatter zone boundary lines. The chatter-free zones are:
4. (nw /γ) H > 4.24 (KH)
5. (nw /γ) L1 < 2.15 (KL1) for γ < 6.67◦
6. (nw /γ) L2 < 1.08 (KL2) for γ > 6.67◦
Figure 16a shows a narrow stable zone in 1.53 < (nw /γ) < 2.15. The setup for this chatter-stable
zone is too risky to use, so it is not considered an area of stable, chatter-free conditions.

129
Inventions 2017, 2, 26

Since the work speed nw is controlled by the regulating wheel speed Nr, it is helpful to convert
the (γ–nw ) charts (Figures 15b and 16b) into a (γ–Nr ) chart, as shown in Figure 17. For practical setup
operations, the range of the center height angle is set to γ = 3◦ to 9◦ . Also, in this case the range of
the speed ratio q (defined as the ratio of the work speed Vw to the grinding speed Vg) for surface
roughness control is set to 1/q = 25 to 150. In Figure 17, three chatter-free stable zones—KH, KL1
and KL2—are shown within the constrained range. The dynamic system stability criterion is satisfied
when a set of (γ, Nr) is selected from the chatter-free stable zones KH, KL1 and KL2.

15

430
Even-lobe Odd-lobe Chatter zone
chatter zone chatter zone

Work speed nw rps


10
n·nw [Hz]

0
0 101.4 180 281.4 360
n·ઞ [°] 0 3 6 9 12
Center-height angle ઞ°
(a) (b)

Figure 16. Dynamic system stability criterion for machine B. (a) Chatter zones on (n·γ–n·nw ) diagram;
(b) Chatter zones on (γ–nw ) chart.

[nw/੘]H

25
Regulating wheel speed Nr rpm

Nru
Speed ratio 1/q=Vg/Vw

KH

Chatter zone

KL1 KL2
Nrl 150

0
3° 6.67° 9°
Center-height angle ੘°

Figure 17. Chatter free zones on γ–Nr chart. KH: High speed chatter-free zone, KL1, KL2: Low speed
chatter-free zones.

5. Modeling to Find the Optimum Setup Conditions that Satisfy the Three Stability Criteria of
Centerless Grinding
The previous section discussed the determination of setup conditions that would satisfy each
individual stability criterion. This section discusses the development of a model for the optimum setup
conditions that will simultaneously satisfy all three stability criteria.
Figure 18 describes the structure of the developed model. As the first step, the sets of (θ, γ)
(blade angle θ, center-height angle γ) satisfying the three individual stability criteria are determined.
To satisfy the work rotation stability criterion, the sets of (θ, γ) are selected from the safe operation zone
shown in Figure 7, analyzed with the varying maximum friction coefficient μr0 of a given regulating

130
Inventions 2017, 2, 26

wheel. Also, the sets of (θ, γ) satisfying both geometrical rounding stability criteria—RW − GRSC and
B − GRSC—are calculated. Then, the sets of (nw , γ) (work speed nw and γ) that satisfy the dynamic
system stability criterion are found. The sets of (nw , γ) are selected from one of the stable chatter-free
zones: KH, KL1 or KL2 (see Figure 17).

Work rotation stability Geometrical rounding stability Dynamic system stability


(90-ɽ-ɴ)/੘=Odd
( ɽ,੘ )
180/੘=Odd
Center-height angle ੘°

( ੘, nw )

Work speed nw (rps)


Absolute Chatter zone
Safe zone
Ȗ

⹄ࡺ⹕⸣
Grinding
( ɽ,੘ )

wheel

䃯ᮤ䓺
Regulating
ਇᶯ
( ੘, nw )

wheel
Blade
ɽ
Blade angle ɽ° Center-height angle ੘

Find the sets of (ɽ, ੘, nw) that satisfy three stability criteria

Determine the optimum set of (ɽ, ੘, nw)

Provide the optimum setup condition (ɽ, CH, Nr)

Figure 18. Structure of the model for determining the optimum setup condition

The second step is to find the sets of (θ, γ, nw ) satisfying all three stability criteria. The third
step is to determine the optimum set from among the population of the (θ, γ, nw ) sets by calculating
the PI (performance index) function based on the process aim (accuracy or productivity). Finally,
the optimum set (θ, γ, nw ) is converted into practical setup conditions (blade angle θ, center height CH,
RW speed Nr) as the outputs of the developed model.
Figure 19 is the flow chart of “Opt-Setup Master”, the developed model. The model requires
a machine operator to provide some input information, as shown in Table 1. All parameters required
for the calculation of the Opt-Setup Master are referenced from the data bank, which stores machine
specifications, machine dynamic characteristics, work part numbers with dimensions, RW friction
characteristics, blade availability, etc. The constraints of the setup parameters are also stored in the
data bank, as shown in Table 2.

Inputs of operational Process priority


DATA BANK information • Grinding accuracy
• Machine specifications • Productivity
• Machine dynamics
• Setup constrains
• Wheel and blade info.
• Work PN and
dimensions Blade availability ɽi Geometrical stability criteria ੘i

Symbols
ɽ: Blade angle Work rotation stability criteria (ɽi, ੘i)
CH: Center-height No
੘: CH angle Safe
Yes
nw: Work speed
Nr: RW speed Dynamic stability criteria (nwi, ੘i) Performance
Index
functions
Priority ranking of setup condition sets (ɽi, ੘i, nwi)

Outputs of optimum setup conditions (ɽi, CHi, Nri)

Figure 19. Flow chart of developed model “Opt-Setup Master”.

131
Inventions 2017, 2, 26

Table 1. Input information and parameters referred to data bank.

Inputs Action Parameters Referred to Data Bank


Machine specifications, Machine dynamic characteristics (Natural
Machine name Select
frequencies, damping ratios)
Workpiece shape Select Cylindrical (CYD), Tapered (TPD), Spherical (SRL), Multi-stepped (STD)
Workpiece part-number Select Dimensions (diameter, length, etc.), profile
GW diameter Measure New, worn, measured
RW diameter Measure New, worn, measured
RW dresser type and dress lead Select Single point dress, Rotary dress, Friction coefficient of RW
Blade availability Select Blade angle θ, blade thickness t

Table 2. Constraints of setup parameters.

Setup Parameters Symbol Unit. Min. Max. Typical


Range of speed ratio 1/q = Vg/Vw - 25 150 100
Range of blade angle θ ◦ 15 45 30
Range of Center-height angle γ ◦ 3 9 6.67
Range of regulating wheel speed Nr rpm 15 100 50
Range of GW diameter Dg mm 375 455 450
Range of RW diameter Dr mm 275 350 345
Range of Workpiece diameter Dw mm 5 100 40
Grinding wheel speed in revolution Ng rpm 1260 2300 1890
Grinding speed Vg m/s 30 45 45

The Opt Setup Master calculates the boundary line of the safe operation zone on the safe operation
chart by using the following relationship between the dressing lead leadr and the maximum friction
coefficient μr0 of the rubber bonded regulating wheel [15].

μr0 = a · leadr + 0.33 (28)

where a is a constant. a = 0.14 and a = 0.02 for a SPD (single-point dresser) and an RD (rotary
dresser), respectively.
The boundary line of the safe operation zone on the (θ, γ) chart can be expressed by:

γc = m · θ c + b (29)

where
m = 3.036 · a · leadr + 1.168 (30)
2
b = 77.06 · ( a · leadr ) − 27.55 · ( a · leadr ) − 21.23 (31)

When θ1 of a point (θ1, γ1) is greater than (γ1–b)/m, the point is located at the right side of the
safe operation zone boundary line and meets the work rotation stability criterion.
The first calculation of the Opt-Setup Master is to find the sets of (θg, γg) that meet the geometrical
rounding stability criteria RW − GRSC and B − GRSC. The next calculation is to determine if the sets
(θg, γg) are located within the safe operation zone under the given dressing conditions of the rubber
bonded regulating wheel. If they are (answer “yes”), the sets (θgw, γgw) satisfy both the geometrical
rounding and the work rotation stability criteria. The final calculation is to find the work speed nw by
using γgw and the chatter stability boundary lines of (nw /γgw). From these calculations, the optimum
sets of (θ, γ, nw ) are discovered.
Then, the performance index (PI) functions that were prepared based on the process aims
are applied to assess the optimum sets of (θ, γ, nw ). The PI functions are summarized in Table 3.
The weighting factors of the PI functions are determined by applying theoretical knowledge,
experimental knowledge and operational skills. PI functions can be updated with newly gained
knowledge and skills. For each process aim (accuracy or productivity), the values of the PI function
for all setup conditions are calculated and these sets are ranked in ascending order from minimum to

132
Inventions 2017, 2, 26

those with greater values. The smaller the PI values, the more suitable the setup conditions. The setup
conditions with the smallest PI function values are defined as the optimum setup conditions.

Table 3. Performance Index (PI) functions.

Weighting Factors
Process Aim PI Function
A B C D E
Accuracy PIa = A × I θ − 27.5 I + B × I γ − 6.67 I + C × I Nr − 50 I + D × FL + E × STYP 0.2 2 0.1 0.01 2
Productivity PIr = A × I θ − 45 I + B × I γ − 5.15 I + C × I Nr − 50 I + D × FL + E × STYP 0.1 2 0.12 0.5 2
θ: Blade angle; γ: Center-height angle; Nr: RW speed in rpm; FL: KL1 = 0, KL2 = 0.5, KH = 1, STYP:
(RW + B) GRSC = −1, RW − GRSC= 0, B − GRSC = +1.

Table 4 shows examples of the Opt-Setup Master outputs. The conditions of the model simulation are:

(1) machine B is applied


(2) the process aim is accuracy-oriented
(3) the work shape and the size are a cylindrical workpiece with 40 mm (D) × 60 mm (L)
(4) the rubber bonded RW is dressed with leadr = 0.5 mm/rev by SPD
(5) the GW diameter is 453 mm
(6) the RW diameter is 350 mm
(7) the available blade angles are θ = 27.5◦ and θ = 40.3◦

Table 4. Examples of outputs from Opt-Setup Master.

Priority Optimum Set Up Conditions Engineering Parameter Stability Parameters


Blade Blade RW Work 1/q Stable RW − Blade −
Ranking Center-Height CH Angle (nw /γ)
Angle Thickness Speed Speed Ratio Zone GRSC GRSC
Nr (90 − θ −
No. θ (◦ ) t (mm) CH (mm) γ (◦ ) nw (rps) Vg/Vw (1/s) KH/KL1/KL2 180/γ
(rpm) β)/γ
1 40.3 20 12.68 41.8 6.68 6.1 25 0.91 KL2 27 7
2 40.3 20 12.66 83.5 6.67 12.2 25 1.83 KL1 27 7
3 40.3 20 13.67 45.1 7.2 6.58 25 0.91 KL2 27 6.5

Conditions: Machine B, Process aim: accuracy, Work: cylindrical Dia.40 × L60 mm.

Ranking No. 1 has the optimum setup conditions, as it has the smallest PI values. The practical
setup parameters are a blade angle θ = 40.3◦ with a thickness of 20 mm, a center height of
CH = 12.68 mm and RW speed of Nr = 41.8 rpm. The center height angle is γ = 6.68◦ , the work
speed is nw = 6.1 rps and the 1/q is 25.
The optimum setup condition set was selected from the safe operating zone. Therefore, it meets
the work rotation stability criterion and ensures safe operations. Further, the optimum setup condition
set was selected from the stable zone KL2 for chatter-free grinding, and thus meets the dynamic system
stability criterion. Also, the optimum setup condition set meets the criteria of both Equations (23)
and (25) (180/6.68 = 27 and (90 − 40.3 − 2.95)/6.68 = 7), so the geometrical rounding stability criteria
are maintained.
Through these means, it is verified that the optimum setup condition set—provided as the outputs
from the Opt-Setup Master—simultaneously satisfies all three stability criteria for centerless grinding.
Table 5 shows the optimum setup conditions as calculated by the Opt-Setup Master for infeed
centerless grinding of cylindrical workpieces of various sizes by two different grinding machines,
A and B. In all cases, γ = 6.67◦ , one of the most preferable γ angles, is chosen. All values representing
RW − GRSC and B − GRSC are odd integers, indicating that all the setup conditions meet the
geometrical rounding stability criteria. Machine B has a greater chatter DSSC index (nw /γ) than
machine A in chatter-stable zones KH, KL1 and KL2. Machine B’s high stiffness creates more extensive
chatter-stable zones than conventional machine A.

133
Inventions 2017, 2, 26

Table 5. The optimum setup conditions provided by the Opt-Setup Master for the grinding of
cylindrical workpieces with different machines A and B.

Work Blade C-H RW Chatter Chatter


Machine RW − GRSC Blade − GRSC
Case Dia.x L. Angle Angle Speed DSSC Stable Zone
No. Nr
mm A/B θ (◦ ) γ (◦ ) (nw /γ) KH/KL1/KL2 180/γ (90 − θ − β)/γ
(rpm)
1 A 27.5 6.67 38.6 3.37 KH 27 9
10 × 20
2 B 27.5 6.67 64.1 5.61 KH 27 9
3 A 27.5 6.67 77.1 3.37 KH 27 9
20 × 30
4 B 27.5 6.67 41.8 1.83 KL1 27 9
5 A 40.3 6.67 14.6 0.43 KL1 27 7
30 × 50
6 B 40.3 6.67 62.7 1.83 KL1 27 7
7 A 40.30 6.67 19.4 0.43 KL1 27 7
40 × 60
8 B 40.30 6.68 41.8 0.91 KL2 27 7
9 A 40.30 6.67 24.3 0.43 KL1 27 7
50 × 70
10 B 40.30 6.68 52.3 0.91 KL2 27 7
11 A 40.30 6.67 29.1 0.43 KL1 27 7
60 × 80
12 B 40.30 6.68 62.7 0.91 KL2 27 7
Condition: GW φ453 mm, Rubber bonded RW φ350 mm, μr0 = 0.4, Available blade θ = 27.5◦ , 40.3◦ .

6. Conclusions
Centerless grinding systems possess some unique features, including their work rotation drive,
loose work holding and regenerative centering mechanisms. However, because of these features,
three fundamental stability issues arise. Many researchers have investigated the issues and provided
useful guidelines for solving the issues.
This paper summarizes the three fundamental stability issues: (1) work rotation stability for safe
operation with no spinners; (2) geometrical rounding stability for better roundness; and (3) dynamic
system stability for chatter-free grinding. It emphasizes the need for an analytical tool that can provide
optimal setup conditions—those conditions that will satisfy all three stability criteria simultaneously.
This paper describes a newly developed analytical tool named Opt-Setup Master, and discusses how
the three stability criteria can be met.
The developed Opt-Setup Master has the following features:

(1) Accepts various shapes of workpiece: cylindrical, tapered, spherical and multi-stepped
(2) Applicable to any centerless grinding machine
(3) Data management via a data bank
(4) Inputs are easy to enter and outputs are readily usable
(5) Designed for operators
(6) Provides scientific parameters for engineers/managers
(7) Finds all setup conditions satisfying the three stability criteria of centerless grinding systems
(8) Outputs the optimum condition based on process aim

Conflicts of Interest: The authors declare no conflict of interest.

References
1. Hashimoto, F.; Gallego, I.; Oliveira, J.F.G.; Barrenetxea, D.; Takahashi, M.; Sakakibara, K.; Stålfelt, H.-O.;
Staadt, G.; Ogawa, K. Advances in Centerless Grinding Technology. CIRP Ann. Manuf. Technol. 2012, 61,
747–770. [CrossRef]
2. Hashimoto, F.; Suzuki, N.; Kanai, A.; Miyashita, M. Critical Range of Set-up Conditions of Centerless
Grinding and Problem of Safe Machining Operation. Jpn. Soc. Precis. Eng. 1982, 48, 996–1001. [CrossRef]
3. Hashimoto, F.; Lahoti, G.; Miyashita, M. Safe Operations and Friction Characteristics of Regulating Wheel in
Centerless Grinding. Ann. CIRP 1998, 47, 281–286. [CrossRef]

134
Inventions 2017, 2, 26

4. Abrasive Industry. Anon, Centerless Grinding Operation; Abrasive Industry: Auckland, New Zealand, 1925;
pp. 102–105.
5. Dall, A.H. Rounding Effect in Centerless Grinding. Mech. Eng. 1946, 68, 325–332.
6. Ogawa, M.; Miyashita, M. On Centerless Grinding (1)—Theoretical Treatise on Rounding Action. Jpn. Soc.
Precis. Eng. 1957, 24, 89–94. [CrossRef]
7. Reeka, D. On the Relationship between the Geometry of the Grinding Gap and the Roundness Error in
Centerless Grinding. Ph.D. Thesis, RWTH, Aachen University, Aachen, Germany, 1967.
8. Rowe, W.B.; Barash, M.M.; Koenigsberger, F. Some Roundness Characteristics of Centerless Grinding. Int. J.
Mach. Tool Des. Res. 1965, 5, 203–215. [CrossRef]
9. Yonetsu, S. Study on Centerless Grinding—First Report. Jpn. Soc. Mech. Eng. 1953, 19, 53–59.
10. Gurney, J.P. An Analysis of Centerless Grinding. Trans. ASME J. Eng. Ind. 1964, 86, 163–174. [CrossRef]
11. Furukawa, Y.; Miyashita, M.; Shiozaki, S. Vibration Analysis and Work Rounding Mechanism in Centerless
Grinding. Int. J. Mach. Tool Des. Res. 1970, 44, 145–175. [CrossRef]
12. Hashimoto, F.; Zhou, S.S.; Lahoti, G.D.; Miyashita, M. Stability Diagram for Chatter Free Centerless Grinding
and Its Application in Machine Development. Ann. CIRP 2000, 49, 225–230. [CrossRef]
13. Gallego, I. Intelligent Centerless Grinding: Global Solution for Process Instabilities and Optimal Cycle
Design. Ann. CIRP 2007, 56, 347–352. [CrossRef]
14. Hashimoto, F.; Lahoti, G.D. Optimization of Set-up Conditions for Stability of the Centerless Grinding
Process. Ann. CIRP 2004, 53, 271–274. [CrossRef]
15. Hashimoto, F. Effects of Friction and Wear Characteristics of Regulating Wheel on Centerless Grinding.
In Proceedings of the SME, Third International Machining and Grinding Conference, Cincinnati, OH, USA,
4–7 October 1999; pp. 1–18.
16. Rowe, W.B.; Koenigsberger, F. The Work-Regenerative Effect in Centerless Grinding. Int. J. Mach. Tool Des.
Res. 1964, 4, 175–187. [CrossRef]
17. Snoeys, R.; Brown, D. Dominating Parameters in Grinding Wheel and Workpiece Regenerative Chatter.
In Proceedings of the 10th International Machine Tool Design and Research Conference, LanPleacaster, UK,
10–14 September 1969; Pergamon Press: Oxford, UK, 1970; pp. 325–348.

© 2017 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

135
MDPI
St. Alban-Anlage 66
4052 Basel
Switzerland
Tel. +41 61 683 77 34
Fax +41 61 302 89 18
www.mdpi.com

Inventions Editorial Office


E-mail: [email protected]
www.mdpi.com/journal/inventions
MDPI
St. Alban-Anlage 66
4052 Basel
Switzerland
Tel: +41 61 683 77 34
Fax: +41 61 302 89 18
www.mdpi.com ISBN 978-3-03842-938-8

You might also like