Sie sind auf Seite 1von 8

Int J Adv Manuf Technol

DOI 10.1007/s00170-013-5198-0

ORIGINAL ARTICLE

Investigation of surface integrity in dry machining


of Inconel 718
Domenico Umbrello

Received: 20 April 2013 / Accepted: 12 July 2013


# Springer-Verlag London 2013

Abstract Machining of advanced aerospace materials have


grown in the recent years although the diffucult-to-machine
characteristics of alloys like titanium or nickel-based alloys
cause higher cutting forces, rapid tool wear, and more heat
generation. Therefore, machining with the use of cooling
lubricants is usually carried out. To reduce the production
costs and to make the processes environmentally safe, the
goal is to move toward dry cutting by eliminating cutting
fluids. This objective can be achieved by using coated tool,
by increasing cutting speed, and by improving the product
performance in term of surface integrity and product quality.
The paper addresses the effects of cutting speed and feed on
the surface integrity during dry machining of Inconel 718
alloy using coated tools. In particular, the influence of the
cutting conditions on surface roughness, affected layer,
microhardness, grain size, and microstructural alteration
was investigated. Results show that cutting conditions have
a significant effect on the parameters related to the surface
integrity of the product affecting its overall performance.
Keywords Machining . Inconel 718 . Surface integrity

1 Introduction
Nickel-based superalloys were created in the 1940s primarily
for gas turbine application due to their long-time strength and
toughness at high temperature and more creep resistance
property than available stainless austenitic steels. Nickel
base superalloys are also used for other applications such
as marine equipment, nuclear reactors, etc. They are used in
D. Umbrello (*)
Department of Mechanical, Energy and Management Engineering,
University of Calabria, Via P. Bucci, Cubo 45/C,
87036 Rende, CS, Italy
e-mail: d.umbrello@unical.it
URL: www.unical.it

these aggressive environments because of their ability to


maintain high resistance to corrosion, mechanical and thermal fatigue, mechanical and thermal shock, creep, and erosion at elevated temperatures [1]. Within the commercially
available nickel-based superalloys, Inconel 718 is the most
frequently used for many applications: aircraft gas turbines,
reciprocating engines, metal processing (e.g., hot work tools
and dies), space vehicles (e.g., aerodynamically heated skins,
rocket engine parts), heat treating equipments, nuclear power
plants, chemical and petrochemical industries, and heat
exchangers.
Although it is the most common nickel superalloy used in
the aerospace industry, some drawbacks are noticed and
within these, the poor machinability is probably the worst.
In fact, machining of nickel-based alloys generate high temperatures at the cutting tool edge, impairing their performance as they are subjected to high compressive stresses
acting on the tool tip. This leads to the plastic deformation of
the tool edge, severe notching and flank wear [1, 2].
Severe wear are also due to the high hot hardness and
strength causes deformation of the cutting tool during machining and the austenitic matrix of nickel alloys which
causes rapid work hardening during machining. The poor
thermal conductivity of nickel-based alloys, raises temperature at the toolworkpiece interface during machining, thus,
it accelerates the undesired tool wear and results in the
shortening of cutting tool life [2, 3].
Therefore, all the cutting parameters, such as tool and
coating materials choice, tool geometry, machining strategy,
cutting speed, feed rate, depth of cut, lubrication, etc., must
be controlled and selected in order to achieve an acceptable
tool life and a correct surface integrity for the machined parts
[4, 5]. Itakura et al. [6] conducted dry turning experiments to
clearly identify the tool wear mechanisms when a commonly
used coated cemented carbide tool cuts Inconel 718. Jindal
et al. [7] studied the relative merits of physical vapor deposition (PVD) TiN, TiCN, and TiAlN coatings on cemented

Int J Adv Manuf Technol

carbide substrate (WC-6 wt% Co alloy) in the turning of


Inconel 718. Prengel et al. [8] performed Inconel 718 turning
tests with a coolant and different PVD-coated carbide cutting
tools.
Furthermore, in order to keep increasing the machining
performance, different assistance methods have been recently developed to replace the conventional process [9]. One
of them presents high-pressure jet assistance (HPJA), which
aims at upgrading conventional machining, using the thermal and mechanical properties of a high-pressure jet of water
or emulsion directed into the cutting zone [1012]. By applying a high-pressure fluid jet to the cutting zone, it is
possible to achieve advantages such as significantly decreased temperature in the cutting zone, prolonged tool life,
and lower forces. These results have also shown improved
surface integrity and better dimensional accuracy of the
produced aeroengine components [13].
As recently described in a literature review proposed by
Ulutan and Ozel [14], a large number of researches have
been carried out in order to investigate and to optimize the
machining process of Inconel 718 alloy in terms of improving quality and surface integrity of the components, increasing productivity and lowering cost. However, the main
works concerning machining Inconel 718 were performed
using cutting fluids even when advanced cooling techniques
like HPJA [1012] or hybrid techniques [15] were involved.
Thus, the effect of dry conditions, especially on the surface
integrity of machined products, was little investigated. Recently, Devillez et al. [16] analyzed the role played by the
suppression of cutting fluid on surface finish and residual
stresses during turning of Inconel 718. They also executed
microhardness measurements at various locations on the
cross section of the machined samples to determine the
machining affected layer by plastic deformation. Their results highlight that the workpiece machined in dry condition
was hardened more than the one machined under wet conditions and the hardened layers were about 250 m beneath the
machined surface. Comparable microhardness values and
gradients were measured by Pawade et al. [17] and by
Ezugwu and Tang [18].
Within the perspective to move toward dry cutting by
eliminating cutting, this paper aims to investigate the effects
of cutting speed and feed on the surface integrity during dry
machining of Inconel 718 alloy using coated tools. The
effects of the cutting conditions on surface roughness, affected layer, grain size variations, and phase changes/modification were investigated.

2 Experimental procedure
Dry orthogonal cutting tests were conducted on Inconel 718
(4299 HV0.05) using a high-speed Mazak computer numerical

control (CNC) turning center and setting a configuration as


schematically indicated in Fig. 1a. In particular, a bar of
347 mm as initial diameter was gently machined in order to
create several disks characterized by thin wall geometry
(10 mm depth and 2 mm thick) spaced 4 mm from each other.
Coated DNMG Sandvik tool (ISO S-DNMG150616) was selected and mounted on a Sandvik DDJNR/L tool holder (providing rake and clearance angles of 6 and 4, respectively) as
shown in Fig. 1b.
In order to avoid effects on the machined surface related
to transient condition (in either feed or speed) due to the
orthogonal configuration, the experiments were executed in
the following sequence:
1. The tool was aligned and brought close to the rotating
workpiece by single-stepping through the CNC program
(i.e., only one block/line of program code is executed
with each press of the button by the operator).
2. All the instruments are set to record mode and the
CNC program is taken out of single-step mode.
3. With the next press of the button the rest of the program
is executed uninterrupted, i.e., the tool enters the workpiece and continues cutting at the prescribed feed rate up
till the prescribed end-of-cut diameter, and then instantaneously retracts at maximum feed. This insures that the
rubbing of the tool against the final machined workpiece
surface is minimal (though it may not be zero) and,
equally importantly, invariant/constant for all the experimental conditions. Further, since the feed rate employed
is in the low range (0.0500.100 mm/rev) according to
the tool makers, and due to the large workpiece diameter
(347 mm at start and 338 at the end) the rpm corresponding to the cutting speeds employed are also low (47
66 rpm), it is possible for the CNC machine's hardware
to change from radially inward feed to outward extraction almost instantaneously for all practical purposes.
Hence, the transient effects were minimal, and this was
confirmed from the force signals recorded during the
cutting.
Disks were machined at varying of three cutting speeds
and three feed rates as illustrated in Table 1; three replications were performed for each test. It is important to underline that the use of these ranges for the cutting parameters
were determined taking into account the tool maker recommendations for typical and industrial machining operations
on nickel-based alloys. The cutting time of each test was 80
90 s in order to reach the mechanical and thermal steadystate conditions.
After machining, samples of 55 mm were sectioned by
wire-EDM, then polished and etched for 35 s using Kalling's
reagent (number 2) to observe microstructural changes (affected
layer and grain size) using a light optical microscope (1,000).
The surface roughness, Ra, of the machined workpiece was

Int J Adv Manuf Technol


Fig. 1 a Scheme of the
orthogonal machining; b
obtained disks from Inconel
718 bar and coated DNMG
Sandvik tool positioned for the
orthogonal machining operation

measured using a Zygo optical white light interferometrybased surface profilometer. The surface and subsurface hardness
values were also measured on a micro hardness indenter Future
Tech F-7. Additionally, the X-ray diffraction (XRD) Bruker
AXS D8 Discover with a quarter Ellurian cradle sample holder
was used for investigating the microstructural phase composition of the machined surface. The X-ray diffractometer used CuK radiation ( =1.54184 , K 1/K 2 =0.5) from a source
operated at 40 kV and 40 mA. Samples were accordingly
positioned at the center of plate into the X-ray goniometric in
order to ensure a correct beam irradiation. The 2 scans were
carried out between 30 and 100 2. The scan increment was
0.05; the corresponding acquisition time was accordingly
varied.

Figure 2 also shows a mapped region called turning


replaces grinding where only cutting speed of 70 m/min
produces comparable Ra values with grinding.
3.2 Microhardness
Figure 3 shows the variation in microhardness values for the
different experimental conditions employed. In particular,
the results highlight that, in all of the investigated cases,
the surface hardness is higher than that of the bulk material.
Also, the value of the ratio HV0.05max/HV0.05initial increases
with the increasing of both cutting speed and feed rate.
Furthermore, higher cutting speed and feed rate allow the
material to reach a deeper hardness variation. The depth
containing hardness values greater than the one in the bulk
material ranges from 6065 m for test ID 1 up to 120
130 m for test ID 9.

3 Experimental results and discussion


3.3 Grain size and affected layer
3.1 Surface roughness
The surface roughness values, Ra, of the machined sample
were measured five times for each test to evaluate the characteristic of the machined surface. As shown in Fig. 2, the
surface roughness measurements for all the test conditions
are always below 0.3 m, which is a very good finish surface
quality. It has been shown that, a smooth surface with better
surface roughness would prevent the initiation of cracks
under cyclic loads [19].

Table 1 Experimental machining test conditions

The structure of each machined surface and subsurface has


been measured by optical microscope (1,000). It was found
that all the examined samples presented a refinement of size
near to the machined surface and beneath it (the initial grain
size in the bulk averages at 18 m). Figure 4 shows the
optical images for tests ID 1 and ID 9, while Fig. 5 shows

Turning replaces grinding

Cutting speed

Feed rate [mm/rev]

0.050
0.075
0.100

50 m/min

60 m/min

70 m/min

ID 1
ID 2
ID 3

ID 4
ID 5
ID 6

ID 7
ID 8
ID 9

Fig. 2 Surface roughness, Ra, on machined samples at varying of


cutting speeds and feed rates

Int J Adv Manuf Technol

Grain size on the bulk material

b
Grain size on the bulk material

c
Grain size on the bulk material

Fig. 5 Grain size evolution near the machined surface and below it: a
50 m/min, b 60 m/min, and c 70 m/min
Fig. 3 Surface and subsurface hardness profiles at varying of cutting
speed: a 50 m/min, b 60 m/min, and c 70 m/min
Fig. 4 Optical images of the
machined surface and subsurface:
a test ID 1 and b test ID 9

affected featureless layer

affected featureless
layer

Int J Adv Manuf Technol

Fig. 6 Thickness of the affected layer measured on machined samples


at varying of cutting speeds and feed rates

Estimation of this layer for all the samples was executed


by Image Pro Plus software; results are reported in Fig. 6. As
observed, the affected featureless layer ranges between 3.5
and 6 m and it increases with both the cutting speed and
feed rate. However, the influence of cutting speed is more
significant when feed higher than 0.075 mm/rev is utilized.
Finally, it is important to underline that depth containing
grains slightly smaller than the one in the bulk material
ranges from 75 m for test ID 1 up to 120130 m for tests
characterized by the higher feed rate (IDs 3, 6, and 9). These
depths are in accordance with those found by analyzing the
microhardness (Fig. 3).

3.4 XRD phase analysis


grain size variation from machined surface to the bulk material for all the investigated tests. Results clearly show that
the cutting conditions influence the final microstructure of
the machined product. In fact, the grain size becomes smaller
when higher cutting speeds and feed rates are utilized.
Moreover, in several tests, the grain size on the machined surface cannot be revealed by optical microscope
even when the largest magnification was used (1,000)
since affected featureless structures appear. The appearance of the featureless layers formed under machining
(Fig. 4b) were similar to the white layers in the machined
surfaces of nickel-based superalloy IN-100 [20], where
significant grain refinement to nanocrystalline level was
found due to dynamic recrystallization.

Figure 7 shows diffraction patterns of Inconel 718 for several


machined samples and for the as received material. In particular, Fig. 7a reports the influence of the feed rate for the cutting
speed of 50 m/min (tests IDs 1, 2 and 3), while Fig. 7b shows
the effect of the feed rate at the higher cutting speed (tests IDs
7, 8, and 9). The X-ray phase analysis on the as received
material shows five peaks at 43.7, 50.9,75.3, 91.3, and
96.4 which, according to Bragg's law, correspond to Ni alloy
in a face-centered cubic (FCC) structure at (111), (200), (220),
(311), and (222) Miller's indices, respectively [21].
According to the Hanalwalt method [22] phase and crystal
structure are defined by the peaks intensity of powder pattern.
In particular, the sequence of the three strongest lines is

(111)

(200)

(111)

(200)
(220)

(311) (222)

Fig. 7 X-ray phase analysis of machined Inconel 718 samples at varying

(220)

(311) (222)

Int J Adv Manuf Technol


Fig. 8 Comparison of XRD
patterns between the as received
sample and the machined sample
at 70 m/min and 0.1 mm/rev
(ID 9)

(111)

(200)

(004)
(200)
(004)

responsible of the different phase and crystal structure. Tests


from ID 1 to ID 6 do not highlight any phase modification,
since their XRD patterns show that the three strongest intensities are ordered as the virgin sample [(111), (200), (311) ].
Then, by applying the Hanalwalt search technique the virgin
material and the machined samples from ID 1 to ID 6 are
characterized by a gamma prime, , structure (PDF 18872).
Gamma prime, the first of the two phases to precipitate during
heat treatment, is a coherent, ordered Ni 3(A1,Ti) face-centered
cubic structure with nickel present on the cube faces and
niobium on the corners of the unit cell.
In contrast, tests carried out at the higher cutting speed (ID
7, 8, and 9) show a different sequence of the three strongest
intensities [(111), (311), (200)], indicating that the structure
on the machined samples is affected by the cutting process.
This evidence is clearer when the highest cutting speed and

(220)

(311)
(222)

feed rate are considered (test ID 9): the presence of precipitating phase (004) at 48.5 2 can be observed (Fig. 8).
Precipitating phase , together with , are responsible
for the heat resistance properties of the matrix gamma ()
phase [23]. Gamma double prime, which nucleates and
coarsens on the particles, is a coherent but misfitting and
ordered metastable body-centered tetragonal Ni3Nb structure
with the nickel atoms sitting on the faces and niobium,
titanium, and aluminum on the corners in the body center
sites. Both and enhance the mechanical properties of
the Inconel 718 alloy by anti-phase boundary strengthening
and coherency strains [2426] although the metastability of
the primary strengthening (, gamma double prime) phase
is typically unacceptable for applications above about
650 C. As a result, other more costly and difficult to process
alloys, like Waspalloy, are used in such applications [27].

Fig. 9 XRD peak and width of Ni alloy in a FCC structure at (111) of both machined and as received samples under a different cutting speeds for the
feed rate of 0.1 mm/rev and for b different feed rates for the cutting speed of 70 m/min

Int J Adv Manuf Technol

Finally, another cardinal aspect to be highlighted from the


XRD analysis is related to the influence of the cutting parameters on the peak relative intensity and its width.
According to Herbert et al. [28], different intensity and width
represent different grain size. As it is clearly seen in Fig. 9,
grain refinements are observed when higher cutting speeds
and feed rates are utilized. Moreover, for the cutting parameters employed in this research, the influence of the feed rate
seems to be more predominant on grain refinements than the
cutting speed. In fact, observing Fig. 9b, the peaks fitting of
the machined samples are almost similar while some differences can be seen when the cutting speed is varied (Fig. 9a),
highlighting that moderate cutting speed does not allow to
reach a high grain refinement.

4 Conclusions
In this paper, an experimental study is proposed for investigating the dry machining of Inconel 718 alloy in terms of surface
integrity indicators (surface roughness, microhardness, affected layer, grain size and phase changes); consequently, the
following conclusion can be drawn:
&

&
&

&

&

Surface roughness in machining Inconel 718 alloy are


comparable with those obtained by finishing processes
(e.g., grinding process), when coated tool are used, and
cutting speeds higher than 70 m/min and low feed rates
are chosen.
Higher cutting speed and feed rate allow the material to
reach a higher surface hardness and a deeper hardness
variation.
All the examined samples presented a refinement of
size and, for certain conditions, the grain size on the
machines surface cannot be revealed by optical microscope since affected featureless structures appear. The
appearance of this layer formed under machining underline that significant grain refinement occurred due
to dynamic recrystallization.
XRD observations highlight that there is a phase change
on the machined surface for tests carried out at 70 m/min.
Also, XRD results show that the peak relative intensity
and its width are influenced by the cutting process parameters. In particular, high grain refinements are observed when higher cutting speeds and feed rates are
utilized. Moreover, the influence of the feed rate seems
to be more predominant on grain refinements than the
cutting speed.
XRD pattern for test ID 9 (70 m/min and 0.1 mm/rev)
shows the presence of precipitating phase which,
together with , are responsible for the heat resistance
properties of the matrix gamma () enhancing the mechanical properties of the Inconel 718 alloy.

Acknowledgments The author gratefully thanks Mr. J. Backus from


Kentucky Geological Survey for his help with the XRD measurements.
The author also acknowledges the undergraduate student Diego Maida
for his contribution with the analysis of the microscope images.

References
1. Ezugwu EO (2005) Key improvements in the machining of difficultto-cut aerospace superalloys. Int J Machine Tools & Manuf 45(12
13):13531367
2. Ezugwu EO, Bonney J, Yamane Y (2003) An overview of the
machinability of aeroengine alloys. J Mater Proc Technol 134(2):
233253
3. Puavec F, Kramar D, Krajnik P, Kopa J (2010) Transition to
sustainable productionpart II: evaluation of sustainable machining technologies. J Cleaner Prod 18(12):12111221
4. Arunachalam RM, Mannan MA, Spowage AC (2004) Surface
integrity when machining age hardened Inconel 718 with coated cutting tools. Int J Machine Tools & Manuf 44(14):1481
1491
5. Thakur DG, Ramamoorthy B, Vijayaraghavan L (2009) Machinability investigation of Inconel 718 in high-speed turning. Int J Adv
Manuf Technol 45:421429
6. Itakura K, Kuroda M, Omokawa H, Itani H, Yamamoto K, Ariura Y
(1999) Wear mechanism of coated cemented carbide tool in coated
tool in cutting of Inconel 718 super-heat resisting alloy. Int Japan
Soc Prec Eng 33(4):326333
7. Jindal PC, Santhanam AT, Schleinkofer U, Shuster AF (1999)
Performance of PVD TiN, TiCN and TiAlN coated cemented
carbide tools in turning. Int J Refract Metals Hard Mater 17(1
3):163170
8. Prengel HG, Jindal PC, Wendt KH, Santhanam AT, Hedge PL,
Penich RM (2001) A new class of high performance PVD coating
for carbide cutting tools. Surf Coat Technol 139(1):2534
9. Shokrani A, Dhokia V, Newmann ST (2012) Environmentally
conscious machining of difficult-to-machine materials with
regard to cutting fluids. Int J Machine Tools & Manuf 57:
83101
10. Lpez de Lacalle LN, Prez-Bilbatua J, Snchez JA, Llorente JI,
Gutirrez A, Albniga J (2000) Using high pressure coolant in the
drilling and turning of low machinability alloys. Int J Adv Manuf
Technol 16:8591
11. Courbon C, Kramar D, Krajnik P, Puavec F, Rech J, Kopa J
(2009) Investigation of machining performance in high-pressure
jet assisted turning of Inconel 718: an experimental study. Int J
Machine Tools & Manuf 49(11):11141125
12. Nandy AK, Gowrishankar MC, Paul S (2009) Some studies on
high-pressure cooling in turning of Ti6Al4V. Int J Machine Tools
& Manuf 49(2):182198
13. Ezugwu EO, Bonney J, Da Silva RB, akir O (2007) Surface
integrity of finished turned Ti6Al4V alloy with PCD tools using
conventional and high pressure coolant supplies. Int J Machine
Tools & Manuf 47(6):884891
14. Ulutan D, Ozel T (2011) Machining induced surface integrity in
titanium and nickel alloys: a review. Int J Machine Tools & Manuf
51(3):250280
15. Feyzi T, Safavi MS (2013) Improving machinability of Inconel 718
with a new hybrid machining technique. Int J Adv Manuf Technol
66:10251030
16. Devillez A, Le Coz G, Dominiak S, Dudzinski D (2011) Dry
machining of Inconel 718, workpiece surface integrity. J Mater
Proc Technol 211:15901598

Int J Adv Manuf Technol


17. Pawade RS, Joshi SS, Brahmankar PK (2008) Effect of machining
parameters and cutting edge geometry on surface of high speed
turned Inconel 718. Int J Machine Tools & Manuf 48:1528
18. Ezugwu EO, Tang SH (1995) Surface abuse when machining cast
iron (G-17) and nickel base superalloy (Inconel 718) with ceramic
tools. J Mater Proc Technol 55:6369
19. McClung RC (2007) A literature survey on the stability and significance of residual stresses during fatigue. Fatigue Frac Eng Mater &
Struct 30:173205
20. Wusatowska-Sarnek AM, Dubiel B, Czyrska-Filemonowicz A,
Bhowal P, Ben Salah N, Klemberg-Sapieha J (2011) Microstructural characterization of the white etching layer in nickel-based
superalloy. Metall Mater Trans A 42:38133825
21. Joint Committee on Powder Diffraction Standards (2010) Powder
Diffraction File, Inorganic Phases, sets: 160. International Centre
for Diffraction Data Inc
22. Joint Committee on Powder Diffraction Standards (2012) Powder
Diffraction File, Hanawalt Search Manual for Experimental Phases,
Inorganic Phases: sets 162. International Centre for Diffraction
Data Inc

23. Collier JP, How Wong S, Tien JK, Phillips JC (1988) The effect of
varying Al, Ti, and Nb content on the phase stability of INCONEL
718. Metall Trans A 19(7):16571666
24. Paulonis DF, Oblak JM, Duvall DS (1969) Precipitation in nickelbase alloy 718. Trans ASM 62:611622
25. Kriege OH, Baris JM (1969) The chemical partitioning of
elements in gamma prime separated from precipitationhardened, high-temperature nickel-base alloys. Trans ASM
62:195200
26. Oblak JM, Paulonis DF, Duvall DS (1974) Coherency strengthening in Ni base alloys hardened by Do22 gamma double prime
precipitates. Metall Trans 5:143153
27. Kushan MC, Uzgur SC, Uzunonat Y, Diltemiz F (2012) ALLVAC
718 Plus superalloy for aircraft engine applications. In: Agarwal
RK (ed) Recent advances in aircraft technology. InTech, Rijeka
(Croatia), pp 7596
28. Herbert C, Axinte D, Hardy M, Brown PD (2012) Investigation into the characteristics on the white layers produced in a
nickel-based superalloy from drilling operations. Mach Sci
Technol 16(1):4052

Das könnte Ihnen auch gefallen