Sie sind auf Seite 1von 10

Characterization of Minerals, Metals, and Materials Edited by A.M. Gokhale, J. Li, T.H.

Okabe TMS (The Minerals, Metals & Materials Society), 2007

COMPARISON BETWEEN TRADITIONAL AND INNOVATIVE STEELS FOR LARGE PLASTIC MOULDS
Maurizio Chiarbonelloa, Donato Firraoa*, Paolo Matteisa, Giovanni M.M. Mortarinoa, Pasquale Russo Spenaa, Giorgio Scavinoa, Graziano Ubertallia, Maria G. Iencob, Gabriella Pellatib, Maria R. Pinascob, Enrica Stagnob, Riccardo Gerosac, Barbara Rivoltac, Agostino Silvestric, Giuseppe Silvac, Adriano Tavascic, Elisa Tatad, Severino Missorid, Roberto Montanarid, Andrea Ghidinie Dip. SMIC, Politecnico di Torino Dip. DCCI, Universit di Genova c Dip. Meccanica, Politecnico di Milano d Universit di Roma Tor Vergata e Lucchini Sidermeccanica S.p.A.
b a

*Corresponding author Keywords: Plastic mold steel, metallography, mechanical property, fracture toughness, fractography. Abstract Moulds for plastic automotive components such as bumpers and dashboards are usually machined from large pre-hardened steel blooms. Due to the blooms size, the heat treatment of the standard 1.2738 steel produces mixed microstructures (continuously varying from surface to core) and a very low fracture toughness, that makes this steel sensible to the defects that may occur in the moulds. In the present work, two alternative steels are investigated: a quenched and tempered microalloyed steel, and a precipitation hardenable steel. Whereas the former can be subjected to the same mould production cycle currently employed for the 1.2738 steel, the latter can be hardened by a subcritical treatment after machining. The microstructures and the ensuing mechanical behavior of steels samples, representative of different positions inside large bloom, have been examined by metallographic techniques, by fracture toughness and tensile tests, and by ensuing fractographic examinations. Introduction Large steel moulds are employed to form automotive components, such as bumpers and dashboards, by the use of thermoplastic polymers, often glass reinforced. The mould service stresses arise from the polymers injection pressure and from the thermal gradients, and can be enhanced by notch effects and by defects (e.g. due to weld bed depositions); moreover, stresses can be significantly raised by abnormal operations (e.g. incomplete extraction of already formed pieces). A mould can produce a few millions of pieces, corresponding to the production run of one car model; thus, fatigue effects should be considered. Wear induced by the reinforced resins flow may be severe and may be an additional cause for crack nucleation and propagation, with the flowing resin infiltrating cracks and acting as a wedge. The moulds are commonly machined from large quenched and tempered blooms, typically 1x1 m section and more than 1 m in length, of the 1.2738 (or 40CrMnNiMo8-6-4 [1]) alloy steel grade. Despite this steels high hardenability, continuously varying microstructures, with decreasing hardness and strength, occur from surface to core in the quenched and tempered

blooms [2, 3]. The blooms resilience and fracture toughness are everywhere unsatisfying (of the order of 10 J and 40 MPam), and remarkably lower than those achieved in samples with similar strength [2, 3], individually treated to yield a complete martensite structure after quench. Moreover, the 1.2738 steel grade is difficult to weld, mainly due to its carbon content (carbon equivalent [4] = 1.16), whereas welding is often used to perform mould shape corrections. Because the moulds shape requires a very deep machining, a production cycle consisting of rough machining, heat treating, and finishing, would allow a more effective quench, but isnt used in practice, because of the increased lead time, and of the additional finishing cost due to the heat-treatment dimensional changes. Several precipitation hardening steels have been proposed as an alternative, with the aim of yielding more uniform microstructures and properties throughout the bloom sections (because their heat treatment is less size-dependent [5]), and of improving the weldability (because of a lower carbon content), in respect to the quenched and tempered grades. The P21 [6, 7] standard grade, for example, contains 0.2% C, 4% Ni, 1.2% Co, and lower amounts of V, Al, Mn, Si, Cr [6]; yet, most grades are proprietary and not disclosed in detail [8]. The solubilization temperature can be subcritical, as for the P21 grade [6, 9] (albeit after an hypercritical annealing [9]), or hypercritical, for some proprietary grades, whereas the aging temperature is always subcritical (e.g. 530 C for the P21 grade [6, 9]), and therefore yields only very limited dimensional variations. The final (service) hardness is usually in the 37-42 HRC [8] range. The final machining can be performed either before or after the aging treatment; the former cycle is normally employed for the P21 grade, and can be employed with other grades if a higher service hardness is sought, but the latter cycle is often preferred for lead-time reasons. In the present work, two new steels, with different compositions and heat treatments, are investigated. Whereas steel 1 belongs to the abovementioned precipitation hardening group, steel 2 is a quenched and tempered, 0.25 % carbon alloy steel, characterized by the addition of 0.6 % Mo and by microalloying with Nb, V, Zr and B (Table I)1. In respect to the traditional quenched and tempered 1.2738 grade, steel 2 is expected to yield an improved weldability, due to its lower carbon content, while keeping a similar strength (due to secondary hardening obtained from the Mo addition) and hardenability (due to the microalloying). Table I. Chemical composition of the examined steels. C Mn Cr Ni Mo Si Nb V Zr Steel 1 0.16 0.68 1.59 0.15 1.36 3.20 1.10 3.18 0.60 0.22 0.28 0.02 0.08 0.12 0.03* Steel 2 0.28 * Estimated.

B 0.0013

Experimental Original Heat Treatment And Bloom Sampling Steel 1 (Precipitation Hardening Steel). The 2400x1500x500 mm forged bloom was subjected to a preliminary hypercritical heat treatment, to a subcritical de-hydrogenization treatment, and to a final hypercritical heat treatment (Figure 1). By comparing the preliminary heat treatment to the steels CCT diagram (Figure 1b, [11]), it can be assumed that: the coarse primary carbides (due to the large molybdenum and vanadium content) were almost completely dissolved during the high-temperature austenitization; the pearlite-ferrite transformation was avoided by the air quenching; fine and homogenously dispersed carbides were re-precipitated in the undercooled austenite matrix during the holding just below the CTT pearlite-ferrite formation temperature
Both steel have been recently proposed by the Lucchini steelwork, with the KeyLOS2001 and KeyLOS2002 trade names, respectively.
1

range and the successive controlled cooling, which caused the matrix to transform to bainite at lower temperatures, no transformation occurring in the final uncontrolled air cooling to room temperature. The de-hydrogenization treatment, although causing (tempering and) aging of the bainitic matrix, probably did not affect the abovementioned carbides. During the final heat treatment, the bloom was austenitized at a lower temperature (the difference being of the order of 50 C), thus the same carbides were at least partially retained; it was then quenched in air, thus yielding a completely bainitic matrix both at surface and at core (on the basis of the same CCT diagram, Figure 1b), which was twice subjected to tempering. The second quench was analyzed with constant-parameter analytical calculations (Figure 1b). The final tempering temperature was low and less than the previously reported [10] aging temperature range, yielding a hardness of about 380 HV. The tested sample was located at a depth of approximately 85 mm from the heat treated surface of the bloom.
1200 Austenitization

Temperature [C]

900 Dehydrogenization 600 Temper (aging) 300

0 0 50 100 150 200 250 300 350

(a)

Time [h]

(b) Figure 1. Steel 1: heat treatment schedule (furnace and air cooling stages represented with continuous and dashed lines, respectively, a); cooling curve in the tested specimen original position superimposed to the CCT diagram [11] (b). Steel 2 (Microalloyed Steel). The heat treatment performed on the steel 2 bloom consists of a dehydrogenization and of an austenitization followed by a water quenching and two tempers, the first one at a higher temperature followed by air cooling, and the second one at a lower temperature with controlled cooling (Figure 2). Three sets of samples, representative of three different positions in the bloom (surface, middle depth and core), have been obtained; each set is composed of samples supposed to be in the same metallurgical conditions for symmetry reasons (Figure 3).

For each samples set, a representative cooling curve was obtained from a constant parameter analytical calculation and compared to the CCT diagram (Figure 4, [12]); these analysis suggests that the microstructure obtained after quenching should be fully martensitic for the surface samples only, whereas the middle depth samples should have a mainly bainitic microstructure, and the core samples could show some pearlite.
1000 900 800 Temperature [C] 700 600 500 400 300 200 100 0 0 100 200 Time [h] 300 400 500 Dehydrogenization Temper Austenitization

Figure 2 Steel 2: heat treatment schedule.

Figure 3 Steel 2: Bloom sampling: original positions of 35 mm thick SENB fracture toughness samples and of 9 mm diameter tensile samples.
Core Surface Middle depth

Figure 4. Steel 2: Cooling curve in the tested specimens original positions superimposed to the CCT diagram [12].

Microstructures The austenitic average grain size was investigated for the two examined steels, by using the circular intercept method [13], after a Nital [14] etch in steel 1, and a Bechet-Beaujard etch [15] in steel 2. The austenitic grain size is on average 130m (mean of 263 intercepts) for steel 1 at the surface position, and 265m (mean of 530 intercepts) for steel 2 at the middle depth position. The actual microstructures were examined at increasing depths from surface to core, after a Nital etch. Steel 1. The microstructure consists mainly of bainite, with small grains size differences among the examined positions, and with small carbide particles, not completely resolved by optical microscopy, probably molybdenum and vanadium carbides (Figure 5).

Figure 5. Bainitic microstructure of steel 2 (2% Nital etch). Steel 2. The as-quenched microstructures must be hypothesized to explain the observed tempered ones. Different cooling speeds at increasing depth during quenching produce a mixture of martensite and bainite near the surface (Figure 6a), and only bainite at core (Figure 6c); both were affected by the following temper. Near the surface the actual microstructure is mainly acicular (Figure 6a), so that orientation relationships can be approximately traced first from austenite grains to martensite and bainite laths formed during quenching, then from martensite laths to still aligned carbides precipitated during tempering. The microstructure coarsens at increasing depths, whereas tempered martensite laths are increasingly substituted by bainite packets, the latter probably affected by carbide coarsening during tempering. (Figure 6b). The core microstructure is completely bainitic (Figure 6c); the higher magnification denotes the presence of some inclusion of zirconium oxide (Figure 6c, higher magnification). Mechanical tests Standard tensile, KIc fracture toughness, and Vickers hardness tests were performed upon the two steels (Table II), as well as FIMEC flat punch indentations [16, 17, 18, 19]. The reported hardness values of each samples set are averages of six indentations. Tensile tests have been performed according to the ISO standard. No apparent necking occurred in the tensile test of the the steel 2 core sample, its ultimate reduction of area being almost uniform. The FIMEC indentation test, performed with a flat cylindrical punch (1 mm diameter) at a specific displacement rate (1.66 m/s), allows to determine the local yield stress as of the 0.2% elastic limit pressure (Figure 7); the present results are higher than the tensile ones (Table II). Fracture toughness tests have been performed according to the ASTM E399 standard; the reported Kq values (Table II) for steel 2 are averages of two tests. In Figure 8a the steel 2 results are compared with the fracture toughness data available for the traditional 1.2738 steel [2, 3].

The load prescribed by the standard for the KIc values calculation always resulted the maximum of a pop-in, and this, for most of the samples, makes the ratio PMAX/PQ (Maximum load / Load used in the calculation) larger than the maximum value indicated in the standard for a valid test (PMAX/PQ<1.1). Only one of the tests, for the core sample of steel 2, is strictly valid according to the standard.

Figure 6. Microstructures of steel 2 (2% Nital etch). Tempered martensite and bainite at the bloom surface (a); coarser tempered martensite and bainite at an intermediate depth (b); bainite at core (c); a zirconium-rich inclusion (c, higher magnification). Table II. Results of the mechanical tests (averages for each set of samples). YS: Yield Strength; UTS: Ultimate Tensile Strength; n: hardening exponent; z: reduction of area. Hardness Fracture Toughness Tension FIMEC KQ PMax/PQ YS UTS n z YS Specimens HV100 MPa MPa % MPa MPam Steel 1 378 66 1.17 868 1156 0.09 35 887 Surface 376 54 1.18 - 1.21 964 1086 0.07 41 989 Steel 2 Middle Depth 377 44 1.23 - 1.20 969 1140 0.06 21 1027 Core 385 44 1.22 - 1.09 996 1157 0.07 \8 1062

Core position (c)

Intermediate position (b)

Surface position (a)

50

50

core middle surface

40

40

Load (Kgf)

30

Load (Kgf) 0,2 0,4 0,6 0,8 Penetration depth (mm) 1,0

30

20

20

10

10

0 0,0

0 0,0

0,2

0,4

0,6

Penetration depth (mm) (a) (b) Figure 7: FIMEC flat indentation load-displacement curves. Steel 1 sample (a) and steel 2 samples from the three examined positions (b).

55
steel 2

50 KQ [Mpam] 45 40 1.2738 35 30 0 100 200

steel 2 core

1.2738 core

300 400 Depth [mm]

500

600

(a) (b) Figure 8. Comparison between fracture toughnesses measured at different depths in steel 1 and in steel 1.2738 (respective bloom cores indicated with dashed lines, a). Load-COD curve of a middle depth steel 2 sample (b). Fractography A SEM analysis has been performed on tensile and SENB samples fracture surfaces. The steel 1 plane-strain fracture surface shows cleavage (Figure 9a) with small microscopically ductile intergranular ruptures (Figure 9b); the latter morphology becomes prevalent in the tensile fracture surface of the same steel, together with some cleavage. The steel 2 plane-strain fracture surfaces are completely brittle, both at surface and at core, and are formed by large cleavage zones and by some small brittle intergranular zones (Figure 10); a similar morphology also occurs on the tensile fracture surfaces obtained from the core position (Figure 11b), whereas some ductile areas appear in the middle depth tension specimen (whose prevalent morphology is still brittle), and become prevalent in the surface tension specimen (Figure 11a).

(a) (b) Figure 9. Steel 1, plane-strain fracture surfaces at the onset of crack propagation; cleavage (a,b) and intergranular rupture (b); image size 187x187 m.

(a) (b) Figure 10. Steel 2, plane-strain fracture surfaces at the onset of crack propagation; 750X, surface sample (a) and 750X, core sample (b); image size125x125 m.

(a) (b) Figure 11. Steel 2, tensile tests samples fracture surfaces; surface sample (a, image size 27x27 m) and core sample (b, image size 47x47 m).

Discussion and Conclusions The microstructure occurring in the steel 2 bloom depends upon the distance from the blooms surface: it consists of tempered martensite and bainite near the surface, showing an increasing fraction of tempered bainite at increasing depths. The core microstructures is bainitic (although the occurrence of pearlite had been hypothesized from the calculated core cooling curve, Figure 4, probably because the constant parameter calculation underestimated the surface heat exchange at the higher temperatures). Therefore, the microstructure of a large steel 2 bloom is overall more homogeneous than that of conventional 1.2738 blooms, the latter showing large pearlite amounts at core [2, 3]. Accordingly, the steel 2 hardness and tensile properties are more homogeneous than those encountered in the 1.2738 steel; in particular its hardness is almost constant, the mean value being 380 HV, and its yield strength and ultimate strength vary only slightly inside the bloom, the former from 960 MPa close the surface to 996 MPa in the core, and the latter from 1085 to 1157 MPa. Nevertheless, the steel 2 tensile tests also show a transition from surface to core toward a more brittle fracture mode, as evidenced by the variation both of the tensile reductions of areas (necking being substantially absent at core) and of the tensile fracture surfaces (being almost completely brittle at core); for comparison, the conventional 1.2738 steel does not show a similar transition, its room temperature tensile behavior being everywhere ductile. The steel 2 fracture toughness is low in comparison to conventional quenched and tempered steels, being about 54 MPam at surface and decreasing to 44 MPam at middle depth and at core. This fracture toughness decrease between the surface and middle depth positions is probably due to the reduction of the amount of tempered martensite with increasing depth. Nevertheless, the steel 2 fracture toughness values are overall somewhat higher in respect to the conventional 1.2738 steel [2, 3], notwithstanding completely brittle plane-strain fracture surfaces of the former steel (opposed to the partially ductile 1.2738 ones). The microstructure of the examined position inside the steel 1 bloom (about 85 mm from the surface) consists almost completely of tempered bainite. By considering this steels CCT diagram and the estimated cooling rates at surface and at core, the microstructure can be hypothesized to be homogeneous and, thus, the examined samples can be considered representative of the whole bloom. The steel 1 ultimate strength is close to the maximum (core) value of steel 2, but the steel 2 yield strength is everywhere higher in comparison to steel 1. The occurrence of the ductile intergranular fracture in steel 1, both in the fracture toughness and in the tensile fracture surfaces, may be related to grain boundary precipitates, probably due to the long heat treatment durations. The single fracture toughness value obtained for the steel 1 bloom is substantially higher in comparison to both steel 2 and steel 1.2738 blooms, and approaches the lower bound of the fracture toughness values previously measured on individually quenched and tempered 1.2738 steel samples. Overall, the steel 2 appears competitive in respect to the conventional 1.2738 steel, but not definitely superior. Among the presently considered steels, steel 1 shows the best combination of toughness and strength, but a definitive conclusion is still to be obtained by tests on further heats produced by the steelmaker, necessary to confirm the hypothesized bloom homogeneity. Acknowledgements Italian Ministry for University and Research, for financial support by research grant PRIN 2005090102.

References 1. ISO 4957:1999, Tool steels (international standard, ISO, 1999). 2. M. Chiarbonello et al., Mechanical Properties of Large Plastic-Mold Steel Blooms, Fracture of nano and engineering materials and structures - Proceedings of the 16th European Conference of Fracture (ECF16), Alexandroupolis, Greece, July 3-7, 2006, ed. E.E. Gdoutos (Dordrecht, Holland: Sprinter, 2006), 433-434. 3. D. Firrao et al., Heat Treatment and Failure Risk of Large Automotive Plastic Molds: a Fracture Mechanics Approach and Property Assessment, 2nd International Conference Heat Treatment and surface engineering in automotive applications - Riva del Garda - 20-22 June 2005. (Milano, Italy: A.I.M, 2005), paper n. 165. 4. N. Bailey et al., Welding Steels Without Hydrogen Cracking (Cambridge, Abington publ: 1973). 5. C.C. Davis, Selection of Materials for Molds for Plastics and Rubbers, Metals Handbook, 9th Ed., Vol. 3, Properties and selection: stainless steels, tool materials and special purpose metals, ed. W.H. Cubberly et al. (Metals Park, OH, USA: A.S.M., 1980) 546-550. 6. A.M. Bayer, T. Basco, L.R. Walton, Wrought tool steels, Metals Handbook - 10th Ed. Vol. 1 - Properties and selection: irons, steels and high performance alloys, ed. J.R. Davis, et al. (Materials Park, OH, USA: A.S.M. Int., 1990) 757-779. 7. G.A. Roberts, R.A. Cary. Tool steels. (Metals Park, OH, USA: ASM, 1980). 8. T. Schade, Steel Selection - Closing the Gap with Offshore Tooling, MoldMaking Tech. Mag., Oct. 2003. 9. P.M. Unterweiser, H.E. Boyer, J.J. Kubbs, eds., Heat Treaterss Guide: Standard Practices and Procedures for Steel (Metals Park, OH: American Society for Metals, 1982). 10. M. Faccoli, A. Ghidini, R. Roberti. A Study of the Strengthening Mechanisms in the Novel Precipitation Hardening Keylos 2001 Steel. Proceedings of 7th international tooling conference - Tooling materials and their applications from research to market - Politecnico di Torino Torino, Italy, 2-5 may 2006, eds. M. Rosso, M. Actis Grande, D. Ugues (Milano, Italy: Ancora, 2006), Vol. 2, 153-161. 11. Keylos 2001 (steel data sheet, Lucchini Sidermeccanica steelwork, 2005). 12. Keylos 2002 (steel data sheet, Lucchini Sidermeccanica steelwork, 2005). 13. ASTM E112-96, Standard Test Methods for Determining Average Grain Size (international standard, ASTM, 1996). 14. ASTM E407-99, Standard Practice for Microetching Metals and Alloys (international standard, ASTM, 1999). 15. S. Bechet, L. Beaujard, Nouveau Ractif pour la Mise en Evidence Micrographique du Grain Austnitique des Aciers Tremps ou Temprs-Revenus, La Revue de Mtallurgie, 52 (1955), 830. 16. R. Mouginot, D. Maugis, Fracture indentation beneath flat and spherical punches, Journal of Materials Science, 20 (1985), 4354-4376. 17. H.Y. Yu, M.A. Imam, B.B. Rath, Study of the deformation behaviour of homogeneous materials by impression tests Journal of Materials Science, 20 (1985), 636-642. 18. P. Gondi, R. Montanari, A. Sili, Small-Scale Nondestructive Stress-Strain and Creep Tests Feasible During Irradiation Journal of Nuclear Materials, 212-215 (1994), 1688-1692. 19. A. Donato et al., A remotely operated FIMEC apparatus for the mechanical characterization of neutron irradiated materials Journal of Materials Science, 258-263 (1998), 446-451.

Das könnte Ihnen auch gefallen