You are on page 1of 9

International Journal of Machine Tools & Manufacture 44 (2004) 12611269 www.elsevier.

com/locate/ijmactool

Tool life determination based on the measurement of wear and tool force ratio variation
S.E. Oraby a,1, D.R. Hayhurst b,
b

Department of Production Engineering, Faculty of Engineering and Technology, Suez-Canal University, Port Said, Egypt Department of Mechanical, Aerospace and Manufacturing Engineering, UMIST, P.O. Box 88, Sackville Street, Manchester M60 1QD, UK Received 19 April 2004; accepted 29 April 2004

Abstract Non-linear regression analysis techniques are used to establish models for wear and tool life determination in terms of the variation of a ratio of force components acting at the tool tip. The ratio of the thrust component of force to the power, or vertical, force component has been used to develop models for (i) its initial value as a function of feed, (ii) wear, and (iii) tool lifetimes. Predictions of the latter model have been compared with the results of experiments, and with predictions of an extended Taylor model. In all cases, good predictive capability of the model has been demonstrated. It is argued that the models are suitable for use in adaptive control strategies for centre lathe turning. # 2004 Elsevier Ltd. All rights reserved.
Keywords: Tool life; Tool wear; Tool force ratio

1. Introduction In the manufacture of metallic components using conventional machine tools, cutter tool wear, its eects on surface nish, and cutter tool failure provide major limitations to the achievement of economic production [1,2]. Techniques for describing tool wear and failure are based either on the use of databases with interpolative and extrapolative techniques, or on the use of explicit mathematical relationships. In both cases, the stochastic nature of tool wear provides an obstacle to the achievement of optimal production conditions. This situation can be improved by using adaptive control (AC) techniques [3,4], which create the possibility to measure and to control the process in situ, and so avoid the need to relate the current machine, workpiece and tool conditions to those conditions under which the existing databases and mathematical equations have been generated.
Corresponding author. Tel.: +44-161-200-3817; fax: +44-161-2004166. E-mail address: d.r.hayhurst@umist.ac.uk (D.R. Hayhurst). 1 On leave. Present address: Department Mechanical Production Technology, College of Technological Studies, PAAET, P.O. Box 42325, Shuwaikh 70654, Kuwait.

Instead of using complex tool databases, it is proposed that models be employed which describe wear; which determine tool life in terms of measurements of in-process variables; and which can, if necessary, be recalibrated during metal cutting. Ideally, the form of the models should be such that the workpiecetool variations, and the variations from machine tool to machine tool, are suciently small not to provide obstacles to their eective use in industry. Some possible approaches are now discussed. It is possible to use a range of process variables to indirectly measure the eects of tool wear. Measurements have been made by many investigators [58]. Theses included the use of on-line electronic equipment/transducers to measure: motor power, spindle torque or even the current drawn by a.c. feed drive servomotors [4,9,10]. Another technique which has gained wide acceptance in sensing wear is the dynamic behaviour of the tooling system [1,1115]. However, one of the most promising techniques for sensing tool wear and breakage involves the measurement and observation of the components of static/dynamic force acting at the tool tip during machining [5,1623]. In a previous study reported by the authors [24] for centre lathe turning, force variation during cutting was found

0890-6955/$ - see front matter # 2004 Elsevier Ltd. All rights reserved. doi:10.1016/j.ijmachtools.2004.04.018

1262

S.E. Oraby, D.R. Hayhurst / International Journal of Machine Tools & Manufacture 44 (2004) 12611269

where U is a function of the process variable, t (min) is the accrued cutting time; and secondly, an equation for tool life T: T Hinit V ;f ;d WFi =Fj ; 4

where Hinit is a function of the initial cutting conditions, and W is a function of the current and initial force ratios. The form and validity of Eqs. (3) and (4) are experimentally investigated in this paper.

3. Components of force and wear at the tool tip In a previous study [24], the variations of the components of force were found to correlate, to diering degrees, with tool wear and breakage. These results are now discussed. Both the feed component of force Fx and the radial component of force Fz were found to be more strongly aected by tool wear than the vertical component of force Fy. This is because Fx and Fz (Fig. 1) are closely related to the sliding and frictional conditions between the tool and workpiece; and, Fy reects the combined eects of Fx and Fz, and their respective friction conditions, in terms of the torque and power required to drive the lathe. At the start of cutting with a virgin tool both of the force components Fx and Fz were found to be almost equal [24]. Subsequently, as wear developed at the cutting edge, the values of Fx and Fz diered and became dependent on how the wear scars were distributed on the cutting edge. Wear scars are usually not evenly distributed on the tool ank and three regions of wear can be distinguished, Fig. 2(a): nose wear NW, over region C; ank wear FW, over region B; and notch wear NCW over region N. The radial component of force Fz has been shown [24] to be most aected by nose wear, NW, while the feed component of force Fx is most inuenced by wear scars in the ank, FW, and notch, NCW, areas. In general, if the wear in one area dominates, then the associated force component is most inuenced. Hence, if an average value of ank wear is used to correlate with the behaviour of Fx and Fz, then an erroneous eect on the individual force components will be predicted. To overcome this and to permit the use of an average value of ank wear, the individual force components are replaced by their resultant or thrust force: Fxz Fx sinv Fz cosv, as shown is Fig. 2(b). The vertical component of force acting at the tool tip, Fy, will be used as the normalising force Fj in Eq. (3). It has been selected for three reasons: (i) It reects the overall torque and power input to the lathe, and hence provides a global measure. (ii) It is the force component which is least sensitive to tool wear.

Fig. 1.

Cutting force components in turning operations.

to correlate well with tool wear and breakage. Both the feed component of force Fx, Fig. 1, and the radial component of force Fz were found to be more aected by tool wear than by the vertical component of force Fy. This suggests that force variation during real-time cutting may provide an accurate and reliable technique to monitor tool wear and breakage. However, one of the drawbacks of such an approach is its non-linear dependence on the control variables, and the inuence of the current toolworkpiece interface conditions. A reformulation of the approach has therefore been proposed [25,26] which seeks to eliminate as many of these aspects as possible. The proposed approach is now outlined.

2. Outline of investigation Mackinnon et al. [26] expressed a component of force Fi acting at the tool tip by Fi j1 jm jti f ai d bi V ci 1 gi W ; 1

where f (mm/rev) is the cutting feed, d (mm) is the depth of cut, V (m/min) is the cutting speed, W (mm) is an average measure of wear, gi is a wear constant, j1 is a lubrication constant, jm is the workpiece material hardness, and jti relates to the tool geometry associated with the force component Fi. In order to reduce the dependency of the relationship between Fi and W upon process conditions, Eq. (1) is normalised with respect to another force measure Fj to yield: Fi jti 1 gi W ai aj bi bj ci cj f d V : Fj jtj 1 gj W 2

In this paper, two new equations are proposed; rstly, a simplied equation for force ratio: Fi UV ;f ;d;t;W ; Fj 3

S.E. Oraby, D.R. Hayhurst / International Journal of Machine Tools & Manufacture 44 (2004) 12611269

1263

Fig. 2. Force system and wear distribution on the cutting edge. (a) Flank wear elements and (b) force system.

(iii) Its magnitude is most sensitive to the workpiece material and to the workpiecetool interface conditions; and, hence its measurement should reect those factors in Eq. (1) which one is seeking to remove through normalisation. Following Colwell [27,28], who rst introduced the force ratio concept, Goforth and Kulkarni [29] and, Choudhury and Kishore [19] have used the same approach to develop models for cutter tool wear and life. They used the ratio of the power, or vertical component of force, Fy, and the feed component of force, Fx, given by Fy =Fx . The novel approach adopted here is to use the ratio of the thrust components of force, Fxz, and the power component of force, Fy, given by Fxz =Fy . The next section outlines the experimental facility used in this research to asses the validity of the proposed models. 4. Experimental procedures and set-up Triple coated (1 lm TiN, 3 lm Al2O3, 5 lm TiC) carbide inserts [Sandvik GC345] were clamped to a threecomponent strain gauge dynamometer [30]. These inserts were used to machine an alloy steel 709M40

(En 19), under dry cutting conditions, using a Colchester Mascot centre lathe. The range of operating parameters used were: cutting speed (V) 50200 m/min, feed (f) 0.060.6 mm/rev; and depth of cut (d) 1.33 mm. Three mutually perpendicular components of the force were measured, Fx, Fy and Fz, as shown in Fig. 1. Tests were periodically interrupted and wear scars were measured at the three dierent locations on the tool ank as shown in Fig. 2(a), namely: nose wear, at the point of maximum projection of the tool into the workpiece; ank wear, at the maximum wear land on the ank; and, notch wear at the point of intersection of the depth of cut, d, with the tool ank. The measure of tool wear used in the paper is the average of the three values measured at the three locations. The selection of the average value of wear at failure Wf 0:25 mm has been carefully made to reect the behaviour over the entire domain of conditions, in which wear at any one of the three zones may predominate. In this way, better correlations have been achieved between model predictions and experimental results than if the ISO recommended value of Wf 0:30 mm had been used. Since the experimental results are to be used to identify explicit mathematical equations, and to t them to the data obtained for subsequent use over the

1264

S.E. Oraby, D.R. Hayhurst / International Journal of Machine Tools & Manufacture 44 (2004) 12611269
in Eq. (5) was also neglected to a good order of d approximation. The validity of using Eq. (5) without the terms in V and d is now examined in this paper using an experimental test programme. From each of the 24 experiments, the initial values of the force components, normalised with respect to Fy0, have been determined and used as input, together with the values of d, f and V, into a numerical regression analysis technique to determine the terms j, ai aj , bi bj and ci cj in Eq. (5). The technique makes use of both forward and backward elimination methods to remove and to insert terms in Eq. (5) according to their statistical signicance in describing the input data. The procedures lead to the elimination of the terms d and V (i.e. bi bj ci cj 0 and the following equations result: bi bj

selected domain of operating parameters, the planning of the 24 experiments has been carried out using the central composite design (CCD) technique.

5. Mathematical model for the initial force ratio Initial values of the force components, denoted by Fx0, Fy0 and Fz0 for the feed, radial and vertical components, respectively, are determined, using a virgin tool, by the intercept of the forcetime history at zero cutting time. The initial force ratios Fi0 =Fj 0 may then be formed using Eq. (2). Since the geometry factors jti and jtj are constants, and the wear W is zero, then Eq. (2) becomes Fi0 =Fj0 jf ai aj d bi bj V ci cj ; 5

where j is a constant. The cutting force, both initially and in the stable cutting region, where no built-up-edge has formed, is principally dependent on the crosssectional area of the material cut; and, is relatively independent of cutting speed [3]. Similar observations have been made by Oraby [31] particularly with regard to the vertical, Fy, and radial, Fz, components of force. In their investigation, the index c of the velocity was found to be relatively constant in the range 0.060.10, ci cj the term in Eq. (5) V 1, since ci cj , and hence it was neglected to a good order of approximation. They also observed a similar eect with regard to the depth of cut d, where bi bj , and hence the term

Fx0 =Fy0 0:193f 0:400 ; Fz0 =Fy0 0:186f 0:370 ; and Fxz0 =Fy0 0:253f 0:405 :

These results are shown graphically in Fig. 3. The feed Fx0 =Fy0 , and the radial Fz0 =Fy0 ratios of the force components are similar in magnitude, whilst the normalised thrust force component Fxz0 =Fy0 is elevated approximately according to the vector addition of Fx0 and Fz0 shown diagrammatically in Fig. 2(b). Since the results of Eq. (6) have been obtained with a greater than 85% factor of determination R2, which corresponds to a 92% correlation factor R, it may be concluded that the normalised force ratios are strongly

Fig. 3.

Relationship between initial force ratio and feed.

S.E. Oraby, D.R. Hayhurst / International Journal of Machine Tools & Manufacture 44 (2004) 12611269

1265

Fig. 4. (a) Experimentally determined variation of force ratios and wear with cutting time for (a) test 5 with V 104 m=min, f 0:2 mm=rev, and d 2:25 mm, (b) test 3 with V 145 m=min, f 0:12 mm=rev, and d 2:5 mm, and (c) test 2 with V 145 m=min, f 0:3 mm=rev, and d 2:0 mm.

1266

S.E. Oraby, D.R. Hayhurst / International Journal of Machine Tools & Manufacture 44 (2004) 12611269

independent of both speed V and depth of cut d. The result demonstrates the signicance of using the force ratio to eliminate the parameters that reect the tool workpiece interfacial conditions.

6. Eect of tool wear on force ratio From the 24 experiments carried out, the results of three have been selected for presentation here. They represent a wide range of cutting conditions and the time variation of wear and of the force ratios Fx =Fy , Fz =Fy and Fxz =Fy are presented in Fig. 4(a)(c). Each is now discussed in turn with a view to identifying the link between the respective forms of wear and the components of force measured. 6.1. Feed dependence In general, Fig. 4 shows that, as cutting proceeds, wear increases causing an increase in the values of the dierent force ratios. The rate of increase of a force ratio depends on the severity of the cutting conditions used. More severe conditions produce a wider wear land and consequently an increase in the force, and in the force component ratios. Fig. 4(a)(c) shows that the dierence between the normalised feed Fx =Fy and the radial Fz =Fy force components decreases as the magnitude of the feed increases. For example, comparison of Fig. 4(b), f 0:12, with Fig. 4(c), f 0:3, shows that as feed is increased, for V 145 and similar values of d 2 2:5, the radial and feed components of force become closer. The same trend is born out by Fig. 4(a), although it is less easy to identify due to dierence in speed. 6.2. Interrelationship between wear and force ratios It was observed in the tests that the wear scars were not uniformly distributed over the three locations shown in Fig. 2. Certain trends have been observed: elevated nose wear aects the radial force component ratio Fz =Fy more than it does the feed force component ratio Fx =Fy . For example, in Fig. 4(a), for Test 5, with V 104, f 0:2, and d 2:25, the radial force ratio Fz =Fy is initially larger than the feed force ratio Fx =Fy , but this situation is reversed after 2.5 min of cutting. Thereafter Fx =Fy > Fz =Fy and wear develops uniformly in the nose and ank areas. For the rst 18 min of cutting, wear scars were observed, using optical examination techniques [31], to be uniformly distributed in the nose and ank regions. After 18 min of cutting, wear in both the ank and nose regions increased, but the growth in ank wear predominated, hence causing a larger increase in Fx =Fy than in Fz =Fy . As cutting con-

tinued nose wear was observed to increase at a higher rate than notch wear, causing the Fz =Fy to increase, on average, at a higher rate than Fx =Fy . The overall eect was that Fz =Fy gradually increased to a value at the end of the test almost equal to that of Fx =Fy . In Fig. 4(b), test results are presented for a higher speed V 145 and a depth of cut d 2:5, but for a lower value of feed, f 0:12. Experimental observations of the wear-time behaviour [31] indicated that notch wear was dominant during the early part of the test resulting in the radial force Fx =Fy being the highest force component throughout the test. After approximately 20 min of cutting, a rapid increase in wear took place in the nose and ank regions, which caused both Fz =Fy and Fx =Fy to increase non-linearly with time. The average wear level shown in the gure closely follows the same behaviour. In Fig. 4(c) test results are presented with V 145, f 0:3 and d 2:0; here, the higher value of feed reduces the dierence between Fx =Fy and Fz =Fy . A uniform distribution of wear observed over the nose region, and over the ank and notch regions, has ensured an approximately equal balance between Fx =Fy and Fz =Fy . It is clear from the foregoing discussion of Fig. 4 that the observations of wear [31] can be linked to the changes in the force component ratios Fx =Fy and Fz =Fy . Examination of all three gures clearly shows that the time variations of both Fx =Fy and Fz =Fy are clearly mirrored in the time variations of wear. In addition, the need is also clear to include both the normalised radial component of force Fz =Fy , and the normalised feed component of force Fx =Fy to obtain an accurate wear-time description. In the next section, the signicance is discussed of the experimental observations of the normalised thrust force component.

6.3. The thrust force component ratio Fxz =Fy and wear In Section 5, it was proposed that the normalised thrust force component Fxz =Fy , cf. Fig. 2(b), be used to describe wear. The normalised thrust force component (Fxz =Fy ) has been plotted in Fig. 4(a)(c) and in all cases is shown to correlate well with wear. Hence, the evidence presented in Fig. 4 clearly supports the use of the normalised thrust force component (Fxz =Fy ) to model wear growth. In the next section, the development of mathematical models is outlined, which describe: (i) the variation of Fxz =Fy as dened by Eq. (3); and (ii) the tool life as dened by Eq. (4).

S.E. Oraby, D.R. Hayhurst / International Journal of Machine Tools & Manufacture 44 (2004) 12611269

1267

7. Mathematical models of tool wear 7.1. Cutter life as a function of wear The discussion in Section 5 regarding experimental observations clearly vindicates the dependency on V, f, d, t and W, where t is the cutting time in minutes, expressed by Eq. (3). To investigate the form of the mathematical relationship which couples these variables with Fxz =Fy , a non-linear multiple regression analysis has been carried out using the 669 data points obtained from the 24 experiments, which has led to the determination of the parameters in Eq. (3): Rxz Fxz =Fy 10:79V 0:306 f 0:393 d 0:188 t0:160 W 0:944 ; 7 with a greater than 81% factor of determination R2, which corresponds to a 90% correlation factor R. This equation is of a form that may be used in an AC strategy for centre lathe turning. When used under these conditions, the force ratio Rxz and the time of cut t may be continuously monitored and, by inversion of Eq. (7), the wear W can be determined using embedded computer processing from the input parameters V, f and d. When the calculated value of W exceeds the value at failure Wf 0:25, then the tool is replaced. It is not always convenient, or expedient, to express cutter life in terms of wear. Very often the time of cutting as expressed in Eq. (4) is a more convenient mea-

sure. In the next section, the mathematical form of Eq. (4) is investigated. 7.2. Cutter life as a function of time A mathematical model which expresses the force ratio (Fxz =Fy ) has been achieved using non-linear regression analysis techniques, similar to those described in Sections 5 and 7, to select those combinations of simple functions which describe the input data to the highest degree of statistical signicance achievable. The model for the cutter tool life of T min was found to be: T 78573:35V 1:712 f 0:714 d 1:107 249:49e78:571Rf Ro ; 8

where Ro Fxz0 =Fy and Rf Fxz =Fy denote, respectively, the original value and the experimental value at failure when W 0:25 mm. Eq. (8) holds for a factor of determination R2 of 0.95, which corresponds to a 98% correlation factor R. The high correlation factor has been achieved since only (24 5) input data points have been made use of out of the 669 data points used to develop Eq. (7). This was achieved by selection of the original condition Ro, and of a single data set from each test, namely, the experimental value of Rf, when the time t is equal to the cutter life T; this occurs when W reaches the failure value Wf 0:25 mm. The value

Table 1 Comparison of experimental and predicted tool lifetime values made using both conventional, Eq. (9), and force ratio models, Eq. (8) Test no. Conditions V (m/min) 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 72 145 145 72 104 104 145 72 72 145 104 104 206 50 104 104 104 104 206 50 104 104 104 104 f (mm/rev) 0.12 0.30 0.12 0.30 0.20 0.20 0.12 0.30 0.12 0.30 0.20 0.20 0.20 0.20 0.60 0.06 0.20 0.20 0.20 0.20 0.60 0.06 0.20 0.20 d (mm) 2.00 2.00 2.50 2.50 2.25 2.25 2.00 2.00 2.50 2.50 2.25 2.25 2.25 2.25 2.25 2.25 3.00 1.50 2.25 2.25 2.25 2.25 3.00 1.50 Tool life values (min) (Wf 0:25 mm) Experimental value 119 13 20 30 36 65 37 79 79 7 50 35 4 140 9 96 24 55 4.5 119 8 65 31 48 Conventional model Eq. (9) 126 10 21 46 39 39 28 62 94 8 39 39 4.5 122 8.5 74 25 58 4.5 122 8.5 74 25 58 Force ratio model Eq. (8) 110 17 25 44 35 46 33 57 86 13 59 36 13 127 18 84 26 56 11 126 16 84 26 56

1268

S.E. Oraby, D.R. Hayhurst / International Journal of Machine Tools & Manufacture 44 (2004) 12611269

Fig. 5. Comparison between experimental lifetimes and values determined using the conventional model (Eq. (9)) and force ratio model (Eq. (8)).

of Rf Rxz being determined from Eq. (7). The rst term on the right-hand side of Eq. (8) gives the tool life as expressed by the cutting conditions, and the second term reects the decrease in life due to the evolution of wear. An advantage of using this approach for adaptive tool life control is that a two-stage approach can be used. In the rst stage, prior to machining, the most appropriate combination of cutting conditions for a given life T can be determined from the rst term on the right-hand side of Eq. (8). In the second stage, once machining has commenced, the entire Eq. (8) is employed, in conjunction with Eq. (7), and the tool is discarded when the right-hand side becomes equal to T. To assess the eectiveness of this approach relative to the well-established second-order tool life model, the following previously calibrated [31], and widely accepted, equation will be used: T expf7:953 7:257lnV 1:121lnV 2 0:371lnf 2 0:810lnd2 0:469lnV lnf g 9

This model depends only on the cutting parameters; and has the form of the extended Taylor equation. A comparison of the tool lifetimes predicted by Eqs. (8) and (9) is made in Table 1 with the experimental data used to establish them. The table lists 24 CCD experiments, together with their respective cutting parameters. Results are graphically illustrated by Fig. 5. It may be seen that in particular tests, the predictions of Eq. (8) are superior for reasons which are now discussed. Firstly, better predictions of the experimental tool lifetimes are made using the force ratio model. This can be observed for test numbers: 4, 5, 6, 7, 8, 9, 11, 12, 14, 16, 18, 23 and 24. Secondly, in situations

where a test has been repeated with the same nominal test parameters, e.g. test numbers: 5, 6, 11 and 12 Eq. (9) predicts a constant tool life of 39 min, whereas Eq. (8) predicts tool lifetimes of 35, 46, 59 and 36 min, respectively, due to the dierent experimental values of Rf at failure; which in all cases are closer to the experimental results. However, in the tests numbered 2, 3, 17, 20 and 22, for which the tool lifetimes T ! 13 min, the predictions made using the force ratio method are in excess of those determined experimentally (non-conservative). The average error being less than 12%. For the remaining tests numbered 10, 13, 15, 19 and 21, for which the tool lifetime T 10 min, the average errors are nonconservative by 130%. For the latter conditions, it is expected that the rst term of Eq. (8) will contribute strongly to the predictive accuracy; and, that if secondorder functions were to be admitted in this term then the corresponding predictions for small T would improve signicantly. It would appear, therefore, that provided the low tool lifetimes behaviour is neglected then the force ratio method Eq. (8) provides a continuous monitoring of tool wear/life from the initial to the nal conditions. Its aspect that provides the exibility necessary for the use of a model in a machine tool AC facility.

8. Conclusions Non-linear regression analysis techniques have been used to establish models for wear, tool life, and initial cutting conditions in terms of force ratios rather than absolute values of force. It has been shown that the

S.E. Oraby, D.R. Hayhurst / International Journal of Machine Tools & Manufacture 44 (2004) 12611269

1269

thrust component of force, when normalised with respect to the power, or vertical, component of force acting on the tool, provides a sensitive measure of nose, ank and notch wear. A model has been developed which describes the initial force ratios as functions of the feed component, other cutting parameters have been shown to be of secondary importance. A model has been developed Eq. (7) which relates an average measure of wear to the cutting parameters, time, and thrust force ratio. A further model has also been established which relates the tool life, to cutting parameters, and to initial and nal/failure thrust force ratios, Eq. (8). Good predictive capability of the model, has been shown by comparison with the predictions of an extended Taylor model, and with the results of experiments. Eq. (8) has the potential for use, with force ratio Rxz measurement, in an AC facility for centre lathe tool management. References
[1] P. Srinivasa Pai, P.K. Ramakrishna Rao, Acoustic emission for tool wear monitoring in face milling, International Journal of Production Research 40 (5) (2002) 1081. [2] B.M.P. Fraticelli, Tool-wear eect compensation under sequential tolerance control, International Journal of Production Research 37 (3) (1999) 639. [3] S.E. Oraby, E.A. Almeshaiei, A. Alaskari, Adaptive control simulation approach based on mathematical model optimization algorithm for rough turning, Kuwait Journal of Science and Engineering (KJSE), An International Journal of Kuwait University 30 (2) (2003) 213. [4] T.-Y. Kim, J. Kim, Adaptive cutting force control for a machining center by using indirect cutting force measurements, International Journal of Machine Tools and Manufacture 30 (8) (1996) 925. [5] Y.D. Liao, Development of a monitoring technique for tool change purpose in turning, Proceedings ot the 26th International Machine Tool Design and Research Conference, 1986, p. 331. [6] N. Constantinides, S. Benneti, An investigation of methods for the on-line estimation of tool wear, International Journal of Machine Tools and Manufacture 27 (1987) 225. [7] K.F. Martin, J.A. Brandon, R.I. Grosvener, A. Owen, A comparison of in-process tool wear measurement methods in turning, Proceedings of the 26th International Journal of Machine Tools Design and Research Conference, 1986. [8] D.E. Dimla Jr., P.M. Lister, N.J. Leighton, Neural network solutions to the tool condition monitoring problem in metal cuttinga critical review of methods, International Journal of Machine Tools and Manufacture 37 (9) (1997) 1219. [9] X. Li, Real-time tool wear condition monitoring in turning, International Journal of Production Research 39 (5) (2001) 981. [10] L. Stein Jerey, H. Kunsoo, Monitoring cutting forces in turning: a model-based approach, Trans ASME Journal of Manufacturing Science and Engineering 124 (1) (2002) 26. [11] L. Ming-Chyuan, J.E. Kannatey, Analysis of sound signal generation due to ank wear in turning, Trans ASME Journal of Manufacturing Science and Engineering 124 (4) (2002) 799.

[12] S.A. Kumar, H.V. Ravindra, Y.G. Srinivasa, In-process tool wear monitoring through time series modelling and pattern recognition, International Journal of Production Research 35 (3) (1997) 739. [13] T. Moriwaki, Application of acoustic emission measurement to sensing of wear and breakage of cutting tool, Bulletin of the Japanese Society of Precision Engineering 17 (3) (1983). [14] M. Lee, C. Thomas, D.G. Wildes, Prospects for In-Process Diagnosis of Metal Cutting by Monitoring vibration Signals, General Electricity Company, Corporate Research Development, Schenetady, NY, 1985. [15] S.E. Oraby, D.R. Hayhurst, Tool wear detection using the system dynamic characteristics, Paper No. 2, The Second International Conference on the Behaviour of Materials in Machining, The Institute of Metals, York, November (1991). [16] S. Das, A.B. Chattopadhyay, Application of the analytic hierarchy process for estimating the state of tool wear, International Journal of Machine Tools and Manufacture 43 (1) (2003) 1. [17] C. Chungchoo, D. Saini, A computer algorithm for ank and crater wear estimation in CNC turning operations, International Journal of Machine Tools and Manufacture 42 (13) (2002) 1465. [18] J.-W. Youn, M.-Y. Yang, A study on the relationships between static/dynamic cutting force components and tool wear, Trans ASME Journal of Manufacturing Science and Engineering 123 (2) (2001) 196. [19] S.K. Choudhury, K.K. Kishore, Tool wear measurement in turning using force ratio, International Journal of Machine Tools and Manufacture 40 (6) (2000) 899. [20] D.E. Dimla, P.M. Lister, On-line metal cutting tool condition monitoring. I: Force and vibration analyses, International Journal of Machine Tools and Manufacture 40 (5) (2000) 739. [21] J.H. Lee, S.J. Lee, One-step-ahead prediction of ank wear using cutting force, International Journal of Machine Tools and Manufacture 39 (11) (1999) 1747. [22] A. Novak, G. Osbahr, Reliability of the cutting force monitoring in FMS installations, Proceedings of the 26th International MTDR Conference, 1986, p. 325. [23] S.E. Oraby, Monitoring of machining processes via force signalspart I: recognition of dierent tool failure forms by spectral analysis, Wear 33 (1995) 133. [24] S.E. Oraby, D.R. Hayhurst, Development of models for tool wear force relationships in metal cutting, International Journal of Mechanical Sciences 33 (2) (1991) 125. [25] I.K. Sung, G.L. Robert, A.G. Ulsoy, Robust machining force control with process compensation, Transactions of the ASME Journal of Manufacturing Science and Engineering 125 (3) (2003) 423. [26] R. Mackinnon, G.E. Wilson, A.J. Wilkinson, Tool condition monitoring using multi-component force measurements, Proceedings of the 26th International MTDR Conference, 1986, p. 317. [27] L.V. Colwell, Methods of sensing the rate of tool wear, Annals of the C.I.R.P. 19 (1) (1971) 647. [28] L.V. Colwell, Tracking tool deterioration by computer (during actual machining), Annals of the C.I.R.P. 23 (1) (1974) 29. [29] R.E. Goforth, N.A. Kulkarni, In-process tool life evaluation by cutting force ratio analysisPart I, Proceedings of the ASME WAM, Boston, November, PED, 9, 1983, p. 67. [30] S.E. Oraby, D.R. Hayhurst, High capacity three-component cutting force dynamometer, International Journal of Machine Tools and Manufacture 30 (4) (1990) 549. [31] S.E. Oraby, Mathematical modelling and in-process monitoring techniques for cutting tools, Ph.D. Thesis, The University of Sheeld (1989).

You might also like