You are on page 1of 11

Journal of Materials Processing Technology 127 (2002) 325±335

Surface integrity of hot work tool steel after high speed


milling-experimental data and empirical models
D.A. Axinte, R.C. Dewes*
Machining Research Group, School of Engineering, University of Birmingham, Edgbaston, Birmingham B15 2TT, UK
Received 9 August 2001; received in revised form 8 May 2002; accepted 8 May 2002

Abstract

High speed milling (HSM) using small diameter solid or indexable insert tungsten carbide end mills, with coatings such as TiAlN is now an
established technology for manufacturing moulds and dies in a range of hardened tool steels. The paper presents experimental results and
corresponding empirical models for workpiece surface integrity (SI) of hardened AISI H13 hot work tool steel, when HSM using solid carbide
ball nose end mills coated with TiAlN. The in¯uence of cutting speed, feed rate and workpiece angle on SI (workpiece surface roughness,
microstructure, microhardness and residual stress) was studied using a full factorial experimental design with two levels of each factor. Where
appropriate, empirical models were proposed in order to predict SI based on cutting parameter inputs. No signi®cant changes in
microstructure and microhardness below the machined surface were found. Within the range of cutting parameters tested, the operating
parameters that gave the lowest workpiece surface roughness did not result in the most compressive residual stress distribution below the
machined surface. A compromise in parameter selection is therefore necessary to achieve desired roughness and integrity.
# 2002 Elsevier Science B.V. All rights reserved.

Keywords: High speed milling; Hot work tool steel; Workpiece surface integrity; Modelling

1. Introduction components directly in their hardened state. The cutting


tools are generally manufactured from micrograin cemented
1.1. High speed milling (HSM) tungsten carbide, employing physical vapour deposited
(PVD) coatings such as TiAlN, TiCrN, TiAlCrN and super-
HSM is currently attracting considerable world-wide lattice products [12,13]. This is in contrast to tools for
interest from mould and die manufacturers, as a means to milling at conventional speeds which generally use carbides
directly machine components in a range of hardened tool with standard grain size and more `standard' PVD coatings
steels (hardness 35±62 HRC) [1]. The process has been used such as TiN and TiCN. The HSM process results in reduced
for a number of decades in the aerospace industry for the production costs and lead times, by over 50% for certain
manufacture of aluminium alloy and composite structural cavity shapes/sizes [14]. Machinability data for the HSM of
components such as wings [2±7]. The same industry has hardened steels is widely available as a result of published
more recently adopted the approach for the machining of information from academic and industrial sources [15±18].
titanium alloy engine parts [8,9] and is considering the This covers hot work tool steels e.g. AISI H13, mould steels
technique for nickel-based superalloy components [9,10,11]. e.g. AISI P20 and cold work tool steels e.g. AISI D2. Data on
Traditionally, the production of moulds and dies generally operating parameters, tool life, workpiece surface rough-
involves conventional machining (end milling, etc.) in the ness, dimensional accuracy, times and costs is readily
annealed (soft) state, followed by heat treatment, electrode obtainable. Cutting speeds are generally in the range
manufacture, electro discharge machining (EDM) and 200±400 m/min, feed rates 0.05±0.2 mm/tooth and axial/
®nish grinding/hand polishing. On the other hand, HSM radial depths of cut 0.2±2 mm. Resulting rotational speeds
of moulds/dies typically involves the use of ball nose and feed rates are commonly 20,000±40,000 rpm and 2000±
and radius end mills, 2±16 mm in diameter, to machine 8000 mm/min, respectively. A fairly recent development is
the use of radius end mills at axial depths of cut up to 20 mm
*
Corresponding author. for high speed roughing of materials such as AISI H13 at
E-mail address: r.c.dewes@bham.ac.uk (R.C. Dewes). 50 HRC [19].

0924-0136/02/$ ± see front matter # 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 9 2 4 - 0 1 3 6 ( 0 2 ) 0 0 2 8 2 - 0
326 D.A. Axinte, R.C. Dewes / Journal of Materials Processing Technology 127 (2002) 325±335

1.2. Workpiece surface integrity (SI) material and no signi®cant white layers or plastic deforma-
tion were evident. Microhardness measurements showed a
Engineered components must satisfy surface texture softening effect up to a depth of approximately 10 mm
requirements (roughness and waviness) and traditionally, beneath the machined surface. Residual stresses were com-
surface roughness (mainly arithmetic average, Ra), has been pressive in most of the experiments with a tendency to reach
used as one of the principal methods to assess quality. zero at about 60 mm below the machined surface. SI data
Certain critical service conditions, however, require that (microstructure, microhardness, residual stress and fatigue
the SI of components is considered, as it will affect service life) when using carbide ball nose end mills to machine AISI
life, safety, reliability and life cycle costs. Manufacturing H13 have also been generated [24]. Microstructural altera-
processes, in particular machining, play a major role in tions when employing a worn cutter were more pronounced
determining the level of SI that can be achieved [20]. with a higher cutting speed, a higher axial/radial depth of
Problems that have been identi®ed include overheating/ cut and with a workpiece inclined at 608 to the tool axis
burning, microcracking, surface irregularities, metallurgical rather than normal to the cutter. A reduction in workpiece
alterations including microstructural distortion, tensile resi- microhardness at the machined surface occurred and
dual stresses and reduction in fatigue life. Such changes residual stresses were predominantly compressive, up to
occur due to thermal/mechanical cycling, microstructural 600 MPa, particularly at 608. The fatigue performance
transformations and mechanical deformation during cutting. of HSM surfaces was superior to those produced using
Thermal effects tend to give tensile residual stresses while EDM. ToÈnshoff et al. [25] analysed SI following high speed
mechanical in¯uences contribute to compressive residual ¯ank milling of high alloyed martensitic stainless steel and
stresses. Another in¯uence on residual stresses is phase reported surface roughness (Ra) <4 mm and tensile residual
transformations. Field et al. [21,22] proposed a `minimum' stress <‡500 MPa when using new and worn 10 mm dia-
SI data set, which involved surface ®nish (roughness), meter carbide end mills, both dry and with ¯ood cutting ¯uid
macrostructure, microstructure and microhardness. They application. The tensile residual stresses were probably due
added residual stress and minimal fatigue testing to give to high temperatures which are likely to occur during ¯ank
a `standard' data set. In addition, their `extended' data set milling.
incorporated in-depth fatigue testing, stress corrosion tests The generally `good' SI results from HSM using ball nose
and a host of mechanical tests (tensile, stress rupture, creep, end mills at low depths of cut has been attributed to high
fracture toughness, etc.). Many destructive and non-destruc- cutting forces of typically 200±1000 N and therefore a high
tive techniques are available to investigate the various SI mechanical effect, particularly when cutting with the centre
effects [22]. Despite publication of these data groups in the of the ball nose end mill and with worn tools, together with
early 1970s, they are still cited in current literature and temperatures <400 8C and a corresponding low thermal
therefore appear to remain valid. In addition to safety critical effect [24].
industries such as aerospace, workpiece SI is important for In contrast to the relatively sparse data on SI of hardened
components such as forging dies, plastic moulds and press steels following HSM, a large number of papers are available
tools, as high thermal and mechanical loads occur during on SI following hard turning of tool steels [26±30].
their use and die/mould service life is critical to the econo- El-Wardany et al. [26,27] reported microstructural, micro-
mics of the processes [23]. hardness and residual stress analysis after high speed turning
of hardened AISI D2 tool steel with PCBN inserts. Micro-
1.3. Workpiece SI following HSM and hard turning structural analysis showed a very thin white layer (<5 mm)
and plastic deformation in the ®rst 10±15 mm from the
The number of published papers that deal with workpiece machined surface. No signi®cant microhardness variations
SI following HSM of hardened steels appears to be relatively below the machined surface were found as a result of
low compared with information on the public domain on tool cutting parameter variation, however the microhardness
life, workpiece surface roughness, etc. Elbestawi et al. [15] was affected by tool wear. Residual stresses on the free
detailed aspects of SI following HSM using 12.7 mm dia- surface were generally tensile but become compressive
meter polycrystalline cubic boron nitride (PCBN) single at depths of approximately 20 mm beneath the machined
tooth ball nose end mills for semi-®nishing and ®nishing surface. In research by AbraÄo and Aspinwall [30], axial
hardened AISI H13. The analysis focused on workpiece fatigue life tests on cylindrical AISI E52100 bearing steel
surface roughness and on metallographic examination of the specimens at a hardness of 60 HRC indicated that fatigue
workpiece below the machined surface. It was found that a performance after turning with a low concentration PCBN
damaged layer (4±6 mm thick) appeared beneath the cut tool was superior to using mixed alumina tools. This was
surface. Dewes et al. [16] presented SI data when using a explained by the higher thermal conductivity of PCBN
32 mm diameter cutter equipped with one PCBN or whisker giving lower interface temperatures than when using mixed
reinforced alumina round insert for HSM the same work- alumina. This contributed to more highly compressive
piece material. The resulting workpiece microstructure was surface residual stresses and therefore improved fatigue
continuous from the machined surface into the body of the life.
D.A. Axinte, R.C. Dewes / Journal of Materials Processing Technology 127 (2002) 325±335 327

The data given in published papers on HSM of hardened Table 1


steels [15±19] indicate that perhaps not surprisingly, opti- Experimental matrix
mum workpiece surface roughness, microstructure, micro- a (8) fz (mm/tooth) v (m/min)
hardness and residual stress distribution are unlikely to be
200 300
achieved for the same combination of cutting parameters.
The in¯uence of each parameter needs to be known, together 0 0.05 Test 1 Test 2
0.10 Test 3 Test 4
with interactions, in order to allow at least, a `pseudo- 60 0.05 Test 5 Test 6
optimisation' of SI. If this is accomplished, models of SI 0.10 Test 7 Test 8
vs. cutting parameters could be used for prediction of output
measures for speci®c experimental setups. Computerised
models of HSM have been produced which predict outputs a ˆ 60 ): horizontal upwards (bfN), tool overhang: 30 mm
such as forces and temperatures [31], together with work- and tool runout <8 mm.
piece surface roughness [32] from operating parameters. The experimental procedure employed a full factorial
One area in which very little work has been reported, design with cutting speed (v), feed per tooth (fz) and work-
however, is the correlation of HSM operating parameters piece angle (a) as factors, each at two levels. New tools were
with subsequent workpiece SI, not least because of the lack used in accordance with the experimental matrix in Table 1
of systematic SI data. Despite carrying out an extensive until maximum ¯ank wear (VBmax) reached 0.1 mm. This
literature search, no published papers could be found on was considered to be a realistic tool wear level for ®nishing
statistical analysis/empirical modelling of SI after HSM of [15]. In addition, the main aim of the work was not to
hardened tool steels. measure tool wear/life, which can be extremely long for this
application (>4 h or >1000 m cut length). Even though tool
1.4. Scope of experimental work and empirical models life was not an output measure, as far as possible, tests were
performed in accordance with ISO 8688-2 [33]. Taking into
The present paper details results, statistical analysis and account previous data, for inducing high compressive resi-
empirical models of SI when HSM hardened AISI H13 hot dual stresses on the workpiece surface [11], a horizontal
work tool steel using solid coated carbide ball nose end upwards (‡bfN) cutting direction was chosen when machin-
mills. The in¯uence of cutting parameters on SI was studied ing at a workpiece angle of 608.
using a full factorial design. Analysis of variance (ANOVA) Workpiece surface roughness was measured using a
was performed to identify the in¯uence of cutting para- portable roughness tester (contact stylus) across the feed
meters on output measures and multiple regression was direction (®ve measurements) every 15 m cut length. When
utilised to obtain empirical models. Based on the models, maximum ¯ank wear reached VBmax ˆ 0:1 mm, samples
examples of response surfaces are presented to allow were cut out from the workpiece using electro discharge wire
prediction of output measures. In the range of parameters machining for subsequent SI analysis (microstructure,
tested no `multicriteria optimum' for SI was found therefore microhardness and residual stress). Taper sections at 308
`pseudo-optimum' solutions are proposed. from the workpiece surface were made across the feed
direction, in order to magnify the surface layer, resulting
in different magni®cations (scales) on the horizontal and
2. Experimental work vertical axes of the photographs. The samples were ground/
polished and etched with 2% Nital (2% nitric acid ‡ 98%
2.1. Machine tool, workpiece material, cutting tools, ethanol, by volume) in order to analyse possible metallur-
cutting parameters and methodology gical changes below the machined surfaces using an optical
microscope equipped with a CCD camera. The same sam-
The research was carried out on a high speed vertical ples were used for assessing microhardness distribution
machining centre with a retro®t high speed spindle below the machined surface. Measurements were carried
(45,000 rpm) equipped with an ISO 20 collet toolholder. out using a Knoop indenter (25 g load) with three measure-
The workpiece material used was hardened hot work tool ments taken at 10 mm intervals between successive readings
steel AISI H13 (hardness 47±49 HRC) in the form of up to 100 mm below machined surface. Adjacent samples
50 mm  80 mm  150 mm blocks. The cutting tools were were used for measuring residual stress distribution below
6 mm diameter ball nose end mills with two teeth and the the machined surface. Measurements were performed
following characteristicsÐmaterial: WC substrate with with an X-ray diffractometer equipped with Cr Ka target
TiAlN coating; helix angle: 308; primary clearance angle: and using the following parametersÐtube voltage/current:
13±158; radial rake angle: 0±38; axial rake angle: 0±38; edge 30 kV/10 mA; measuring angle, c: 0, 15, 30 and 458;
preparation: sharp. Cutting parameters held constant were: aperture: 2 mm  4 mm; ISO-inclination method. Measure-
axial depth of cut, ap: 0.2 mm; radial depth of cut, ae: ments were taken along the tool feed direction on the
0.2 mm; cutting method: down milling; cutting environment: free surface and below the machined surface at 10 mm
compressed air at 8 bar; cutting direction (when operating at intervals between successive readings. The residual stress
328 D.A. Axinte, R.C. Dewes / Journal of Materials Processing Technology 127 (2002) 325±335

measurements were stopped when at least two consecutive beginning of tests followed by a sharp decrease and a
measurements were in the interval 0  25 MPa. tendency to level out when the tool reached the stable wear
Data were subsequently analysed in order to ®nd the stage after approximately 100 m length cut.
in¯uence of cutting parameters on SI. ANOVA (signi®cant Empirical modelling was carried out in two stages. Initi-
at the 5% level, i.e. p coef®cient <0.05) was used to identify ally, statistical analysis of the variation of Ra with length cut
the in¯uences of variable cutting parameters and cutting (L) was performed separately for each test. Simple regres-
length/depth below machined surface on the variation of sion analysis on the effect of Ra on L, using a range of
output measures. Main effects and interaction effects were different functions gave a polynomial function as best ®t, see
plotted. As the output tables from the ANOVA contained Eq. (1) where j, k, m, n and p are coef®cients that depend on
extensive information these are summarised in this paper. the variable cutting parameters used.
Factors and/or interactions which were found to be signi-
®cant on variation of the output measures, were used as Ra;Ti ˆ jL ‡ kL0:5 ‡ mL1:5 ‡ nL2 ‡ pL3 (1)
inputs for multiple polynomial regression in order to obtain Ra;Ti represents the workpiece surface roughness evaluated
empirical models. These were also used for plotting surface during a single test, denoted Ti, where i is the test number.
responses. Commercial statistical software (Minitab 13) was The model is valid for all sets of test conditions and the
utilised throughout the work. coef®cients j, k, m, n and p have different values for each
test. The form of the polynomial formula (Eq. (1)) was used
2.2. Results, empirical modelling and discussion as an input for the second stage. This involved statistical
analysis of all test results in order to obtain a general Ra
2.2.1. Workpiece surface roughness model in which all factors (cutting parameters) and length
The workpiece surface roughness measurements (Ra cut are used as inputs. Since the ANOVA proved that all
across the feed direction) are presented in Fig. 1a for test factors and interactions (®rst, second and third level) were
numbers 1±4 (workpiece angle, a ˆ 0 ) and in Fig. 1b for signi®cant for Ra variation (5% level), the model should
test numbers 5±8 (workpiece angle, a ˆ 60 ). Ra values ideally contain all factors and interactions. As the relation-
were in the range 0.36±2.18 mm which is within what is ship between Ra and length cut followed a polynomial
generally required for components such as forging dies and function (Eq. (1)), however, interactions between length
plastic/rubber moulds. The repeatability of the measure- cut and cutting parameters would have been extremely
ments (standard deviation based on ®ve measurements dif®cult to incorporate into the model. For this reason,
divided by the mean value) was found to be in the range the model only takes into account interactions between
of 2±7%. This was considered satisfactory for generating the cutting parameters. Multiple regression for Ra gave a
empirical models. It should be noted that the characteristic polynomial model as presented in Eq. (2) where workpiece
variation in Ra with length cut (an increase at the beginning angle, a is expressed in radians.
of each test followed by a reduction and a tendency to level
out when the test reached a stable wear stage after approxi- Ra ˆ a1 ‡ b1 v ‡ c1 fz ‡ d1 a ‡ e1 vfz ‡ g1 va ‡ h1 fz a
mately 100 m length cut) was the same throughout the ‡ i1 vfz a ‡ j1 L ‡ k1 L0:5 ‡ m1 L1:5 ‡ n1 L2 ‡ p1 L3 (2)
experimental programme.
In order to analyse the in¯uence of cutting parameters on For the coef®cients a1 ˆ 6:642, b1 ˆ 0:008263, c1 ˆ 12:042,
Ra variation, ANOVA was performed, see Table 2. Fig. 2 d1 ˆ 0:9257, e1 ˆ 0:06784, g1 ˆ 0:008679, h1 ˆ
shows the corresponding main effects plots for Ra. In such 7:297, i1 ˆ 0:06144, j1 ˆ 1:4216, k1 ˆ 5:239, m1 ˆ
diagrams, the higher the gradient or length of each line, the 0:16546, n1 ˆ 0:007462 and p1 ˆ 6:01  10 6 , a coef®-
larger the effect of the parameter variation. It can be seen cient of determination, r 2 ˆ 0:91 (corresponding coef®cient
that the mean value for Ra increased when cutting speed of correlation r ˆ 0:95) and a standard error, S:E: ˆ
increased. This is contrary to what would normally be 0:12 mm were obtained. All of these values were obtained
expected (higher cutting speeds generally give lower rough- from the statistical analysis software using the least-squares
ness due to avoidance of built up edge effects) [20]. As no method. It should be noted that the model is only valid in the
built up edge was seen on the tools, the small increase was range of cutting parameters tested and for lengths cut, L
thought to be due to increased unbalance of the tool at the between 0 and 200 m (for a ˆ 0 ) or 0±150 m (for a ˆ 60 ).
higher cutting speed, together with possibly higher cutting Eq. (2) allows the prediction of Ra, from the combination
forces leading to vibrational effects. The small decrease in of cutting parameters and length cut. Based on Eq. (2), an
the mean value of Ra at the higher feed per tooth was thought example response surface for Ra using the empirical model
to be due to higher tool vibration caused by the low feed per for a ˆ 0 and v ˆ 300 m/min is shown in Fig. 3.
tooth. A much larger in¯uence on mean Ra was workpiece
angle. The low mean value at 608 was due to avoidance of 2.2.2. Microstructure and microhardness
cutting in the tool centre and corresponding smearing of No white layers or other microstructural alterations were
workpiece material giving high Ra (which occurs at a ˆ 0 ). observed on all samples even at high magni®cations
The main effects plot indicates that mean Ra increased at the (1000). With higher values of cutting speed and/or feed
D.A. Axinte, R.C. Dewes / Journal of Materials Processing Technology 127 (2002) 325±335 329

Fig. 1. Workpiece surface roughness, Ra vs. length cut when workpiece angle: (a) a ˆ 0 ; (b) aˆ608. Error bars represent 95% confidence intervals based on
five repeated measurements.

per tooth, it is expected that higher temperatures will be to the rubbing effect of the centre of the tool with a 08
generated and that a greater quantity of heat will be trans- workpiece angle (cutting speed zero) can be seen. In Fig. 4b
ferred into the workpiece, thus increasing the possibility of the feed marks are a lot less pronounced.
heat affected zones such as white layers. Microstructural Microhardness measurements were in the range 613±
photographs taken from samples obtained after tests 4 and 8 667 HK0.025. In order to assess the in¯uence of variable
with worn tools (both under `worst case' conditions of cutting parameters and depth below the machined surface (x)
higher levels of cutting speed and feed per tooth) are shown on microhardness variation with worn tools, ANOVA was
in Fig. 4a and b, respectively. In Fig. 4a, the feed marks due performed and the results are summarised in Table 3. The
330 D.A. Axinte, R.C. Dewes / Journal of Materials Processing Technology 127 (2002) 325±335

Table 2 data indicate no signi®cant changes in microhardness with


Outline of ANOVA for Ra depth below the machined surface or signi®cant differences
Inputs 2 levelsÐcutting speed, v; feed per tooth, between tests. These are in line with published results
fz; workpiece angle, a [16,25] which also show no substantial alterations in micro-
8 levelsÐlength cut, L (levels equivalent to 15 m structure or microhardness following HSM of hardened tool
length cut intervals)
steel.
Output Ra (five repeated measurements at each length
cut interval)
2.2.3. Residual stress
Results Based on cut-off p < 0:05 for input significance The experimental results for residual stress measured
output variationÐrejecting null hypothesis along the tool feed direction are presented in Fig. 5a for
Main effects v, fz, a, L were found to be significant
Interactions All interactions (first, second and third level)
test numbers 1±4 (workpiece angle, a ˆ 0 ) and in Fig. 5b
were found to be significant for test numbers 5±8 (workpiece angle, a ˆ 60 ). Stresses
were highly compressive, up to 760 MPa, which is in

Fig. 2. Main effects plot for workpiece surface roughness, Ra.

Fig. 3. Example 2D response surface for workpiece surface roughness, Ra, when cutting speed, v ˆ 300 m/min and workpiece angle, a ˆ 0 .
D.A. Axinte, R.C. Dewes / Journal of Materials Processing Technology 127 (2002) 325±335 331

Fig. 4. Microstructure of workpiece from: (a) test no. 4 (v ˆ 300 m/min; fz ˆ 0:1 mm/tooth; a ˆ 0 ); (b) test no. 8 (v ˆ 300 m/min; fz ˆ 0:1 mm/tooth;
a ˆ 608).

agreement with most other HSM data [16,24] and can be i.e. 6, including x ˆ 0 mm as the ®rst level. Repeatability of
explained by the likely overriding mechanical rather than the measurements changed with depth below the machined
thermal effect. Both graphs follow the same tendency of surface. On the free surface, variation between replicated
decreasing compressive stress with increasing depth beneath data was 15±20%, probably due to the relatively high surface
the machined surface. In order to explain the in¯uence of roughness which scattered the X-ray beam. Below the free
cutting parameters and depth below the machined surface (x) surface, where the roughness of the surface was lower due to
on residual stress variation, ANOVA was performed, the the electrochemical machining used to remove the surface
results of which are outlined in Table 4. Since the residual layer, the variation between replicated readings was in the
stress recovered to 0  25 MPa at different depths for the
various tests, the number of levels of depth below the
Table 4
machined surface used in the ANOVA was the minimum Outline of ANOVA for residual stress below the machined surface

Table 3 Inputs 2 levelsÐcutting speed, v; feed per tooth, fz; work-


Outline of ANOVA for microhardness below the machined surface piece angle, a
6 levelsÐdepth below machined surface, x (levels
Inputs 2 levelsÐcutting speed, v; feed per tooth, fz; workpiece angle, a equivalent to 10 mm increment)
10 levelsÐdepth below machined surface, x (levels equivalent
Output Residual stress (s) with three repeated measurements
to 10 mm intervals)
at each depth below the machined surface interval
Output Microhardness (HK0.025) with three repeated measurements at
Results Based on cut-off p < 0:05 for input significance
each depth below the machined surface interval
output variationÐrejecting null hypothesis
Results Based on cut-off p < 0:05 for input significance output Main effects v, fz, a, x were found to be significant
variationÐrejecting null hypothesis: none of the inputs (v, fz, Interactions First order: vx, fza, fzx; second order: fzax; third order:
a or x) were found to be significant vfz ax were found to be significant
332 D.A. Axinte, R.C. Dewes / Journal of Materials Processing Technology 127 (2002) 325±335

Fig. 5. Residual stress vs. depth below surface when workpiece angle: (a) a ˆ 0 ; (b) a ˆ 60 . Error bars represent 1 standard deviation based on three
repeated measurements.

range of 6±15%. At depths where the residual stresses had As they were probably relatively low (<400 8C), tempera-
recovered to 0  25 MPa, high percentage variation was tures are likely to have had only a small in¯uence on the
found (>20%). Subsequent comments and analysis relate workpiece surface, a tendency which was also seen in the
only to mean values of residual stress. microstructure and microhardness data analysis. The empiri-
Main effect plots for residual stress are presented in Fig. 6. cal modelling for residual stress was performed in two
It can be seen that increasing cutting speed and feed per stages, in a similar fashion to that for workpiece surface
tooth caused the mean level of compressive stress to roughness, with depth below the machined surface (x)
decrease, probably due to a higher thermal effect on the replacing length cut in the analysis. Simple regression
workpiece surface. At the higher workpiece angle, the mean was performed on the residual stress vs. x data. It was found
compressive stress decreased slightly due to the absence of that the best ®t was a square root function of x which was
the rubbing effect which occurs at the centre of the ball nose used as the input into the second stage regression analysis.
end mill and tends to induce compressive residual stresses. Based on the ANOVA results, multiple regression was
D.A. Axinte, R.C. Dewes / Journal of Materials Processing Technology 127 (2002) 325±335 333

Fig. 6. Main effects plot for residual stress.

Fig. 7. Example 2D response surface for residual stress when cutting speed, v ˆ 200 m/min and workpiece angle, a ˆ 0 .

performed, see Eq. (3), where a ˆ 1175, b ˆ 1:2, c ˆ 5190, Based on Eq. (3), an example residual stress response sur-
d ˆ 165, e ˆ 115, f ˆ 0:0895, g ˆ 1518, h ˆ 403, face for a ˆ 0 and v ˆ 200 m/min is shown in Fig. 7.
j ˆ 125 and k ˆ 0:732 are coef®cients and workpiece
angle, a is expressed in radians. For this case, a coef®cient
of determination, r2 ˆ 0:94 (corresponding coef®cient of 3. Conclusions
correlation, r ˆ 0:97) and a standard error, S:E: ˆ 29 MPa
were obtained. All of these values were obtained from the  Workpiece surface roughness (Ra) values obtained
statistical analysis software using the least-squares method. throughout the tests were 0.36±2.18 mm. Ra values
The model is only valid for the range of cutting parameters increased when cutting speed increased and feed per tooth
tested and for a depth below the machined surface, x of decreased due to higher process instability. Ra values
between 0 and 50 mm decreased with the 608 workpiece angle due to the absence
of the rubbing effect caused by the centre of the tool at 08.
s ˆ a ‡ bv ‡ cfz ‡ da ‡ ex0:5 ‡ fvx0:5 ‡ gfz a ‡ hfz x0:5
 No significant white layers or other heat affected zones
‡ jfz ax0:5 ‡ kvfz ax0:5 (3) were found below the machined surface in all the tests.
334 D.A. Axinte, R.C. Dewes / Journal of Materials Processing Technology 127 (2002) 325±335

None of the variable parameters affected microhardness Advances in Materials and Processing Technologies, Dublin, Ireland,
significantly. 1993, pp. 21±32.
[6] B. Ashley, High speed machiningÐthe kindest cut of all, aluminium
 Highly compressive surface residual stresses occurred up industry, October/November 1992.
to 760 MPa. Significant factors were cutting speed, feed [7] P. Dennis, Machining fibre-reinforced plastics, Ind. Diam. Rev. 6
per tooth, workpiece angle, depth below the machined (1991) 288±291.
surface and some of their interactions. Increasing cutting [8] K. Maekawa, I. Ohshima, Y. Nakano, High speed end milling of
Ti±6Al±6V±2Sn titanium alloy, advancement of intelligent production,
speed and feed per tooth caused the compressive stress to
in: Proceedings of the Seventh International Conference on Pro-
decrease due to a likely increase in thermal effect on the duction/Precision Engineering and Fourth International Conference
machined surface. With the higher workpiece angle, the on High Technology, Chiba, Japan, 1994, pp. 431±436.
compressive stress was lower due to the absence of the [9] S. Derrien, J. Vigneau, High speed milling of difficult to machine
rubbing (mechanical) effect caused by the centre of ball alloys, in: Proceedings of the First French and German Conference
nose end mill. on High Speed Machining, Metz, France, 1997, pp. 284±294.
[10] E.O. Ezugwu, I.R. Pashby, High speed milling of nickel-based
 `Best' values of workpiece surface roughness and residual superalloys, J. Mater. Process. Technol. 33 (4) (1992) 429±437.
stress were not achieved with the same cutting conditions [11] E.-G. Ng, D.W. Lee, A.R.C. Sharman, R.C. Dewes, D.K. Aspinwall,
therefore, a `pseudo-optimal' solution is proposed. `Best' J. Vigneau, High speed ball nose end milling of Inconel 718, Ann.
values of workpiece surface roughness and residual stress CIRP 41 (1) (2000) 41±47.
were not achieved with the same cutting conditions there- [12] Application GuideÐDie and Mould Making, Sandvik Coromant, 1999.
[13] W.-D. MuÈnz, I.J. Smith, L.A. Donohue, A.P. Deeming, P. Halstead,
fore, a `pseudo-optimal' solution is proposed, whereby if PVD coated cemented carbide tools designed for dry high speed
low workpiece surface roughness (Ra) is the main goal, cutting applications, in: Proceedings of the First French and German
the cutting parameters of test numbers 3 or 7 should be Conference on High Speed Machining, Metz, France, 1997, pp. 173±
chosen, while if high compressive residual stress is the 183.
[14] M. Page, A milling force to be reckoned with, Metalworking Prod.
main requirement, the parameters used in test numbers 1
138 (4) (1994) 23±36.
or 5 should be selected. It should be underlined that both [15] M.A. Elbestawi, L. Chen, C.E. Becze, T.I. El-Wardany, High-speed
solutions have the advantage of keeping the same cutting milling of dies and moulds in their hardened state, Ann. CIRP 46 (1)
parameters when the workpiece angle changes which is (1997) 57±62.
what happens in reality when machining moulds and dies. [16] R.C. Dewes, D.K. Aspinwall, S.J. Dipple, S. Schoen, G. Thielemann,
Tool wear and surface integrity observations during the high speed
milling of hardened die steel, in: Proceedings of the International
Conference and Exhibition on Design and Production of Dies and
Acknowledgements Molds, Istanbul, Turkey, 1997, pp. 131±138.
[17] P. FallboÈhmer, C.A. RodrõÂguez, T. O È zel, T. Altan, High-speed
We would like to thank Prof. A.A. Ball, Head of the machining of cast iron and alloy steels for die and mold
School of Manufacturing and Mechanical Engineering and manufacturing, J. Mater. Process. Technol. 98 (2000) 104±115.
[18] R.C. Dewes, D.K. Aspinwall, A review of ultra high speed milling of
Prof. M.H. Loretto, Director of the IRC in Materials for High hardened steels, J. Mater. Process. Technol. 69 (1997) 1±17.
Performance Applications for provision of facilities. We are [19] N. Narutaki, Hard metal cutting technology, in: Proceedings of the
grateful to Mr. J. Wedderburn who performed much of the Sharing Tomorrow's Technology Today Conference, Machine Tool
experimental work. Thanks also goes to the Engineering and Technologies Association (MTTA), London/Cambridge/Warwick,
Physical Sciences Research Council (EPSRC), Mr. C. UK, 1997.
[20] Machining Data Handbook, third ed., Vol. 2, Machinability Data
Wheelhouse of United Engineering Forgings Ltd., Mr. R. Center, Metcut Research Associates Inc., Cincinnati, OH, 1980.
Hirons and Mr. A. Smith of Sandvik Coromant, UK, Mr. D. ISBN 0936974001.
Edwards of Matsuura Machinery plc and Mr. M. Fleming of [21] M. Field, J.F. Kahles, J.T. Cammett, A review of measuring methods
De Beers Industrial Diamonds Ltd. for surface integrity, Ann. CIRP 21 (2) (1972) 219±238.
[22] M. Field, J.F. Kahles, Review of surface integrity of machined
components, Ann. CIRP 20 (2) (1971) 153±163.
[23] V. Vazquez, M. Knoerr, T. Altan, R. Shivpuri, Determination of
References fatigue properties of die steels for hot forging, Trans. North Am.
Manuf. Res. Inst. SME, NAMRC XXIV (1996) 155±160.
[1] H. Schulz (Ed.), Hochgeschwindigkeitsbearbeitung (High-speed [24] R.C. Dewes, High speed machining of hardened ferrous alloys, Ph.D.
Machining), Carl Hanser, Munich, 1996. ISBN 3446187960. Thesis (commercially confidential), University of Birmingham, UK,
[2] F.J. McGee, High-speed machining-study: methods for aluminium 1997.
workpieces, Am. Mach. 123 (3) (1979) 121±126. [25] H.K. ToÈnshoff, C. Gey, H. Tùnnessen, K. Sùrby, High speed flank
[3] J. Scherer, The technology of high-speed milling of various milling of Greek Ascoloy, the effect of cooling lubrication on tool
aluminium alloys, in: Tagungshandbuch (Handbook) 4, Darmstadter wear, cutting forces and surface integrity, in: Proceedings of the Second
Fertigungstechniches Symposium (Manufacturing Engineering Sym- International Seminar on Improving Machine Tool Performance,
posium), TH Darmstadt, Germany, 1989. Nantes-La Baule, France, 2000.
[4] F.J. McGee, High-speed machining of aluminium alloys, in: Proceed- [26] T.I. El-Wardany, H.A. Kishawy, M.A. Elbestawi, Surface integrity of
ings of the High Speed Machining, the Winter Annual Meeting of the die material in high speed hard machining. Part 1: micrographical
ASME, Vol. 12, New Orleans, LA, PED, 1984, pp. 205±216. analysis, Trans. ASME, J. Manuf. Sci. Eng. 122 (11) (2000) 620±631.
[5] I. Nieminen, J. Paro, V. Kauppinen, High-speed milling of advanced [27] T.I. El-Wardany, H.A. Kishawy, M.A. Elbestawi, Surface integrity of
materials, in: Proceedings of the International Conference on die material in high speed hard machining. Part 2: microhardness
D.A. Axinte, R.C. Dewes / Journal of Materials Processing Technology 127 (2002) 325±335 335

variations and residual stresses, Trans. ASME, J. Manuf. Sci. Eng. [31] I.F. Dagiloke, A. Kaldos, S. Douglas, B. Mills, High speed
122 (11) (2000) 632±641. machining: an approach to process analysis, J. Mater. Process.
[28] H.K. ToÈnshoff, C. Arendt, R. Ben Amor, Cutting hardened steel, Technol. 54 (1995) 82±87.
Ann. CIRP 49 (2) (2000) 547±566. [32] Y.-H. Tsai, J.C. Chen, S.-J. Lou, An in-process surface recognition
[29] A.S. Sadat, Effect of high cutting speed on surface integrity of AISI system based on neural networks in end milling cutting operations,
4340 steel during turning, Mater. Sci. Technol. 6 (1990) 371±375. Int. J. Mach. Tools Manuf. 39 (4) (1999) 583±605.
[30] A.M. AbraÄo, D.K. Aspinwall, The surface integrity of turned and [33] Tool life in millingÐPart 2: end milling, 1989. ISO 8688-2 (E).
ground hardened bearing steel, Wear 196 (1996) 279±284.

You might also like